Developmental Psychopathology, Risk, Resilience, and Intervention [3 ed.] 9781119125532, 9781118120934

Examine the latest research merging nature and nurture in pathological development Developmental Psychopathology is a fo

207 81 10MB

English Pages 2413 Year 2016

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Developmental Psychopathology, Risk, Resilience, and Intervention [3 ed.]
 9781119125532, 9781118120934

Citation preview

Table of Contents Title Page Copyright Dedication Preface to Developmental Psychopathology, Third Edition References Contributors Chapter 1: Childhood Adversity and Adult Physical Health Defining Childhood Adversity Defining Health Outcomes Childhood Adversity and Later Disease: Epidemiological Evidence Conceptual Models Linking Childhood Adversity to Adult Physical Health Biological Intermediaries Linking Early Adversity to Adult Physical Health Concluding Comments and Directions for Future Research References Chapter 2: Community Violence Exposure and Developmental Psychopathology Introduction: Violence Is Common but Problematic Violence Perpetration and Exposure Are Important but Poorly Defined Developmental Psychopathology Constructs Rates and Patterns of Exposure to Community Violence What Are the Mental Health and Behavioral Effects of Community Violence? Risk and Protective Factors Associated with Community Violence Exposure and Effects Theories about How Community Violence Exposure Causes or Increases Individual Risk for Psychopathology Neurodevelopmental Processes that Might Be Implicated in Community Violence Exposure Interventions for Community Violence Exposure Advancing Knowledge, Practice, and Policies Related to Community Violence Exposure Conclusion References Chapter 3: Social Support and Developmental Psychopathology

Defining Social Support Social Support and Stress Social Support in Relationships and Social Networks Giving and Receiving Social Support in a Cultural Context Interim Conclusions Social Support and Developmental Psychopathology Conclusions and Future Directions References Chapter 4: Poverty and the Development of Psychopathology Rationale and Background Poverty and Psychopathology Mediators and Moderators of Poverty's Effects on Psychopathology Practice and Policy Future Directions References Chapter 5: Determinants of Parenting Introduction Parenting and Parents Some Methodological Considerations and Future Directions Determinants of Parenting Determinants in the Parent Determinants in the Child Determinants in the Context Translational Implications Conclusions References Chapter 6: Resilience in Development: Progress and Transformation Introduction Historical Perspectives Concepts and Models Methods Psychological Promotive and Protective Processes Child-in-Context Resilience Processes Neurobiological Promotive and Protective Processes

Translational Research and Implications Controversies Old and New Progress and Future Directions References Chapter 7: Vulnerability and Resiliency of African American Youth: Revelations and Challenges to Theory and Research Historical and Contextual Framing of Vulnerability and Resiliency Human Vulnerability and Resiliency Methodological Considerations Translational Practices Conclusion: Reframing the Future References Chapter 8: Social Inequalities and the Road to Allostatic Load: From Vulnerability to Resilience Introduction Social Inequalities Influencing Allostatic Load Innovative Biochemical, Neurological, and Cogntive Perspectives Conclusions References Chapter 9: Competence and Psychopathology in Development Defining Competence and Psychopathology: Historical Legacies and Theoretical Work Relations between Competence and Psychopathology Impairment and Mental Disorder: Systems, Measures, and Research Support Cascade Models Interventions to Promote Competence and Reduce Psychopathology Conclusions and Future Directions References Chapter 10: The Development of Coping: Implications for Psychopathology and Resilience Goal of the Chapter Transactional Perspectives: Coping as Individual Differences in Appraisal and Coping Processes and Resources Normative Developmental Perspectives: Coping as a Set of Basic Adaptive Processes that Are Reorganized with Age Developmental Systems Perspectives: Coping as Part of Developmental Cascades toward Psychopathology and Resilience

Future Research and Translation of Research into Action Summary and Conclusion References Chapter 11: Temperament and Developmental Psychopathology Introduction What Is Temperament? The Structure of Temperament Temperament and Personality The Measurement of Temperament Temperament and Developmental Psychopathology Temperament and Psychopathology: Review of the Research Literature Temperament × Temperament Interactions Neurobiology Linking Temperament and Developmental Psychopathology Final Remarks: Summary and Future Directions Translational Implications Conclusion References Chapter 12: Interparental Conflict and Child Adjustment Introduction and History Interparental Conflict and Child Adjustment Problems Theory Explaining the Link between Interparental Conflict and Child Adjustment Child Attributes and Family/Community Factors Contributing to Individual Differences Translational Implications Directions for Future Research Conclusion References Chapter 13: Relational Aggression: A Developmental Psychopathology Perspective Introduction Defining Relational Aggression Developmental Change in Relational Aggression Aggression and Gender Risk Factors: Biobehavioral Processes Risk Factors: Cognitive and Emotional Processes Risk Factors: Social Processes

Developmental Outcomes: Maladaptive Correlates Developmental Outcomes: Potential Positive Correlates Cultural Perspectives Relational Aggression Interventions Developmental Psychopathology Perspectives Future Directions Conclusion References Chapter 14: Culture, Peer Relationships, and Developmental Psychopathology Theoretical Perspectives on Culture, Peer Relationships, and Adaptive and Maladaptive Development Socioemotional Functioning and Problems in Peer Settings Across Cultures Culture and Parental Attitudes Culture, Peer Evaluation, and the Regulatory Function of Peer Interactions Children's Peer Experiences and Adaptive and Maladaptive Outcomes: The Role of Culture Conclusions, Practical Implications, and Future Directions References Chapter 15: Classroom Processes and Teacher–Student Interaction: Integrations with a Developmental Psychopathology Perspective Considering the Intersection of Education and Human Development Educational Demands and Opportunities that Shape Student Experience and Outcomes Conceptualizing and Measuring Teacher–Student Classroom Interactions Psychological and Contextual Factors Related to Qualities of Teachers' Interactions with Students Activating the Classroom as a Setting for Development Future Directions: Deepening Knowledge on the Links between Classroom Processes and Development of Children and Youth The Developmental Science of Applied Practice in Schools Individualization of Classroom Process Effects on Teachers and Students: Biological Sensitivity to Experience References Chapter 16: Advances in Prevention Science: A Developmental Psychopathology Perspective Introduction

Selective Review of Programmatic Preventive Interventions Current Directions in Conceptualizing Complexity Methodological Considerations and Challenges Translational Research Future Directions and Recommendations References Chapter 17: Culturally Adapted Preventive Interventions for Children and Adolescents Why Cultural Adaptation in Developmental Psychopathology Definition of Culture Historical Context of Cultural Adaptation Research The Rationale for Cultural Adaptation: Guidelines toward a more Systematic and Targeted Approach The Content of Cultural Adaptation: Surface Structure, Deep Structure, and Core Components Process Models of Cultural Adaptation: Systematic Application of Best Practices Parent Training Interventions Youth Risk Prevention Programs Prevention of Anxiety and Mood Disorders Health-Focused Interventions Conclusions and Future Research Directions References Chapter 18: The Effects of Early Psychosocial Deprivation on Brain and Behavioral Development: Findings from the Bucharest Early Intervention Project Précis History of Institutional Care in Romania Effects of Institutionalization on Neuropsychological Functioning Experimental Design and Methodology Summary Implications of the BEIP for Developmental Psychopathology Conclusions References Chapter 19: Preventing Sensitization and Kindling-like Progression in the Recurrent Mood Disorders Introduction Types of Sensitization Mechanisms in the Recurrent Affective Disorders and Their

Cross-Reactivity Neurobiological Commonalities in Stress, Episode, and Substance Abuse Sensitization Therapeutic Implications Modulation of Sensitization Mechanisms Lessons from the Neurobiological Mechanisms Involved in Kindling Progression Neurobiological Correlates of Illness Progression in the Recurrent Affective Disorders More Malignant and Progressive Course of Bipolar Disorder in the United States Compared with Some European Countries Public Health Implications Conclusions References Chapter 20: Mental Health Stigma: Theory, Developmental Issues, and Research Priorities Defining Stigma and Stigmatization Theoretical Perspectives on Stigma Historical Perspectives on Stigma: Change and Cyclicity Empirical and Cultural Evidence for Stigma Developmental Themes Toward an Integrative Model of the Social Stigma of Mental Illness Key Research Directions Strategies for Overcoming Stigma References Author Index Subject Index End User License Agreement

List of Illustrations Chapter 4: Poverty and the Development of Psychopathology Figure 4.1 Conceptual model of SES and parenting. Chapter 5: Determinants of Parenting Figure 5.1 Mothers with College Education. Figure 5.2 Unmarried Mothers. Figure 5.3 Mothers Ages 18–64 with a Young Child at Home. Figure 5.4 Age of First-Time Mothers.

Figure 5.5 Mothers' expected number of children. Figure 5.6 Births to Unmarried Mothers. Figure 5.7 Hours Spent Working Outside the Home per Week by Mothers. Figure 5.8 Hours Spent on Housework per Week by Mothers. Figure 5.9 Mothers with Children Younger than 18 in the Labor Force. Figure 5.10 Households with the Mother as the Primary or Sole Source of Income. Figure 5.11 Stay-at-Home Mothers in Poverty. Figure 5.12 Percentage of Single-Parent Families with Children under 18. Chapter 6: Resilience in Development: Progress and Transformation Figure 6.1 Resilience Pathways After Acute Trauma. Model of resilience pathways after acute trauma. The gray zone represents normal or typical adaptive function. Path A represents stress resistance. Path B illustrates breakdown and recovery. Path C represents posttraumatic growth. Figure 6.2 Resilience Pathways after Chronic Adversity. Model of resilience pathways showing a period of chronic exposure to severe adversity followed by a period when conditions improve. Path A shows declining adaptive function during chronic adversity followed by recovery when conditions improve. Path B shows a normalization pattern when conditions improve. Figure 6.3 Risk or Promoter Gradient. A cumulative risk or promoter gradient illustrating how adaptive function falls as accumulating risk factors rise or promotive factors fall. Stars represent individuals who are functioning well even though their risk levels are high. Figure 6.4 Risk and Promoter Model. Basic “main effects” model of risk and promotive factors that influence an adaptive outcome of interest. This model includes an intervention with a promoter effect. Figure 6.5 Mediated Risk Model with Intervention. Risk affects an intermediate variable that in turn changes adaptive outcome. This model includes an intervention intended to indirectly improve the outcome by mitigating risk effects on the mediator or by supporting the mediator. Figure 6.6 Moderated Risk Model. Moderated risk model showing one independent moderator that mitigates or enhances the effects of the risk (adversity) variable and one risk-activated moderator that is triggered by risk or adversity and then buffers or exacerbates the effects of the risk/adversity. Figure 6.7 Differential Sensitivity Model. Moderator model of differential sensitivity to experience showing one “differential sensitivity” moderator that enhances or attenuates the impact of both bad (adversity) or good (advantage) experiences and a

second “vantage sensitivity” moderator that moderates the impact only of good (advantage) experiences. Vulnerability (a moderator of adverse experiences) is not shown. Chapter 8: Social Inequalities and the Road to Allostatic Load: From Vulnerability to Resilience Figure 8.1 (A) A transdisciplinary framework in relation to other research frameworks (adapted from Juster et al., 2011). Multidisciplinarity approaches combine knowledge from different disciplines additively. Interdisciplinary approaches involve the appreciation of disciplinary overlap and the integration of discipline-based constructs into higher-order concepts encompassing several disciplines. Transdisciplinarity emerges as a higher-order framework that integrates all knowledge into new concepts and perspectives that encompass complex interactions reaching across disciplinary boundaries. This approach is consistent with perspectives espoused in the field of developmental psychopathology and is best measured by triangulating methods. Adapted from “A transdisciplinary perspective of chronic stress in relation to psychopathology throughout life span development,” Development and Psychopathology, 2011, Volume 23, Issue 3, p. 761, Cambridge University Press. (B) Allostatic load is an exemplar of trandisciplinarism (Juster et al., 2011). Aspects of AL are illustrated here as topic triangles forming a holistic triangle according to diverse disciplines; namely, biological, psychological, sociological, behavioral, and spiritual. As reviewed in Juster, McEwen, and Lupien (2010), each of these life domains has been empirically substantiated in relation to multisystemic biomarkers representing AL that are consequently associated with clinical outcomes. In relation to Figure 8.A, AL is best understood from a transdisciplinary perspective whereby complex interactions among risk factors and protective factors can be identified. Chapter 10: The Development of Coping: Implications for Psychopathology and Resilience Figure 10.1 Coping Depicted as a Transactional Process of Appraising and Dealing with Demands. Figure 10.2 Four Models of the Role of Coping in the Processes that Connect Stress to Psychopathology, as a (1) Moderator; (2) Mediator; (3) Mechanism; and (4) Reciprocal Process. Figure 10.3 Depiction of the Coping System as a Set of Fundamental Adaptive Processes Used to Detect, Respond to, and Learn from Encounters with Potential Challenges, Threats, and Dangers. Figure 10.4 Integrative Multilevel Conceptualization of Coping as a Set of Interrelated Processes that Function on the (1) Neurophysiological, (2) Psychological, (3) Action, (4) Interpersonal, and (5) Societal Levels. Figure 10.5 Underlying Neurophysiological Factors and Overarching Socialization Factors that Contribute to the Differential Development of Maladaptive Coping and Increase the Risk of Behavior Problems and Psychopathology.

Chapter 12: Interparental Conflict and Child Adjustment Figure 12.1 Organizational Framework for Mechanisms and Pathways by which Interparental Conflict and Violence are Theorized to Influence Child Adjustment. Chapter 13: Relational Aggression: A Developmental Psychopathology Perspective Figure 13.1 The Reformulated Social Information Processing Model. Figure 13.2 Gender-Linked Model of Aggression Subtypes. Chapter 18: The Effects of Early Psychosocial Deprivation on Brain and Behavioral Development: Findings from the Bucharest Early Intervention Project Figure 18.1 Example of Congregate Care in One Romanian Institution. Figure 18.2 Illustration of the initial recruitment of children into the study and over time, their placement status at age 8 years. Chapter 19: Preventing Sensitization and Kindling-like Progression in the Recurrent Mood Disorders Figure 19.1 Parallelisms in Kindling of Seizures and Affective Episodes. Figure 19.2 Cross-Sensitization among Stressors, Drugs of Abuse, and Episodes. Figure 19.3 Postulate: The ratio of pathological versus adaptive changes accounts for cyclic emerge of episodes and tolerance development. Figure 19.4 More Vulnerability Factors and Bipolar Illness Adversity in the United States Compared to Europe. Figure 19.5 Bipolar Disorder in the United States Compared to the Netherlands and Germany. Figure 19.6 Bipolar Disorder Stage Progression in Childhood-Onset versus AdultOnset Illness: Schematic Illustration.

List of Tables Chapter 10: The Development of Coping: Implications for Psychopathology and Resilience Table 10.1 Three Perspectives on How the Study of the Development of Coping Can Contribute to Research on Developmental Psychopathology and Resilience Table 10.2 Six Dimensions of Parenting Chapter 11: Temperament and Developmental Psychopathology Table 11.1 Dimensions of contemporary developmental models of temperament Chapter 17: Culturally Adapted Preventive Interventions for Children and Adolescents Table 17.1 Models of Cultural Adaptation by Date of First Publication

Chapter 18: The Effects of Early Psychosocial Deprivation on Brain and Behavioral Development: Findings from the Bucharest Early Intervention Project Table 18.1 Attachment-Related Measures Table 18.9 Illustration of the Social Emotional Development Battery that Was Deployed at Different Ages Chapter 19: Preventing Sensitization and Kindling-like Progression in the Recurrent Mood Disorders Table 19.1 N-acetylcysteine (NAC) Dampens Conditioned (Cued) Glutamate Release in the Nucleus Accumbens: Is this a Common Mechanism for Suppressing Habit-Based Psychopathologies? Table 19.2 Common Effects of Stress-, Episode-, and Cocaine-Induced Sensitization on BDNF in Brain Systems Mediating Habit and Representational Memory Table 19.3 Differential Pharmacology as a Function of the Stage of Amygdala-Kindled Seizure Evolution Table 19.4 Does Drug Effectiveness Vary as a Function of Stage of Bipolar Illness Evolution? Table 19.5 Neurobiological Correlates of Number of Mood Episodes or Duration of Illness: Evidence of Illness Progression Table 19.6 Consequences of Childhood Adversity in Adulthood

DEVELOPMENTAL PSYCHOPATHOLOGY THIRD EDITION Volume Four: Risk, Resilience, and Intervention Editor DANTE CICCHETTI

This book is printed on acid-free paper. Copyright © 2016 by John Wiley & Sons, Inc. All rights reserved. Published by John Wiley & Sons, Inc., Hoboken, New Jersey. Published simultaneously in Canada. No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning, or otherwise, except as permitted under Section 107 or 108 of the 1976 United States Copyright Act, without either the prior written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, (978) 7508400, fax (978) 646-8600, or on the web at www.copyright.com. Requests to the Publisher for permission should be addressed to the Permissions Department, John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, (201) 748-6011, fax (201) 7486008. Limit of Liability/Disclaimer of Warranty: While the publisher and author have used their best efforts in preparing this book, they make no representations or warranties with respect to the accuracy or completeness of the contents of this book and specifically disclaim any implied warranties of merchantability or fitness for a particular purpose. No warranty may be created or extended by sales representatives or written sales materials. The advice and strategies contained herein may not be suitable for your situation. You should consult with a professional where appropriate. Neither the publisher nor author shall be liable for any loss of profit or any other commercial damages, including but not limited to special, incidental, consequential, or other damages. This publication is designed to provide accurate and authoritative information in regard to the subject matter covered. It is sold with the understanding that the publisher is not engaged in rendering professional services. If legal, accounting, medical, psychological or any other expert assistance is required, the services of a competent professional person should be sought. Designations used by companies to distinguish their products are often claimed as trademarks. In all instances where John Wiley & Sons, Inc. is aware of a claim, the product names appear in initial capital or all capital letters. Readers, however, should contact the appropriate companies for more complete information regarding trademarks and registration. For general information on our other products and services please contact our Customer Care Department within the United States at (800) 762-2974, outside the United States at (317) 572-3993 or fax (317) 572-4002. Wiley publishes in a variety of print and electronic formats and by print-on-demand. Some material included with standard print versions of this book may not be included in e-books or in print-on-demand. If this book refers to media such as a CD or DVD that is not included in the version you purchased, you may download this material at http://booksupport.wiley.com. For more information about Wiley products, visit www.wiley.com. Library of Congress Cataloging-in-Publication Data: Developmental psychopathology / editor, Dante Cicchetti. – Third edition. pages cm Includes index. ISBN 978-1-118-12087-3 (volume 1 : cloth : alk. paper) – ISBN 978-1-118-12091-0 (volume 2 : alk. paper) – ISBN 978-1-11812092-7 (volume 3 : alk. paper) – ISBN 978-1-118-12093-4 (volume 4 : alk. paper) 1. Mental illness–Etiology. 2. Developmental psychology. 3. Mental illness–Risk factors. 4. Adjustment (Psychology) I. Cicchetti, Dante. RC454.4.D483 2016 616.89–dc23

These volumes are dedicated to Marianne Gerschel in recognition of her great vision and staunch support of the field of developmental psychopathology.

Preface to Developmental Psychopathology, Third Edition A decade has passed since the second edition of Developmental Psychopathology was published. The two prior editions (Cicchetti & Cohen, 1995, 2006) have been very influential in the growth of the field of developmental psychopathology. The volumes have been highly cited in the literature and have served as an important resource for developmental scientists and prevention and intervention researchers alike. In the present third edition, we have expanded from the three volumes contained in the second edition to four volumes. The increased number of volumes in this current edition reflects the continued knowledge gains that have occurred in the field over the past decade. A not insignificant contributor to this growth can be found in the very principles of the discipline (Cicchetti, 1984, 1990, 1993; Cicchetti & Sroufe, 2000; Cicchetti & Toth, 1991, 2009; Rutter & Sroufe, 2000; Sroufe & Rutter, 1984). Theorists, researchers, and prevention scientists in the field of developmental psychopathology adhere to a life span framework to elucidate the numerous processes and mechanisms that can contribute to the development of mental disorders in high-risk individuals as well as those operative in individuals who already have manifested psychological disturbances or who have averted such disorders despite their high-risk status (Cicchetti, 1993; Masten, 2014; Rutter, 1986, 1987, 2012). Not only is knowledge of normal genetic, neurobiological, physiological, hormonal, psychological, and social processes very helpful for understanding, preventing, and treating psychopathology, but also deviations from and distortions of normal development that are seen in pathological processes indicate in innovative ways how normal development may be better investigated and understood. Similarly, information obtained from investigations of experiments of nature, highrisk conditions, and psychopathology can augment the comprehension of normal development (Cicchetti, 1984, 1990, 1993; Rutter, 1986; Rutter & Garmezy, 1983; Sroufe, 1990; Weiss, 1969). Another factor that has expedited growth within the field of developmental psychopathology has been its ability to incorporate knowledge from diverse disciplines and to encourage interdisciplinary and translational research (Cicchetti & Gunnar, 2009; Cicchetti & Toth, 2006). In keeping with its integrative focus, contributions to developmental psychopathology have come from many disciplines of the biological and social sciences. A wide array of content areas, scientific disciplines, and methodologies has been germane. Risk and protective factors and processes have been identified and validated at multiple levels of analysis and in multiple domains. The increased emphasis on a multilevel, dynamic systems approach to psychopathology and resilience, the increased attention paid to gene–environment interplay in the development of psychopathology and resilience, and the application of a multiple levels of analysis developmental perspective to mental illnesses that have traditionally been examined

nondevelopmentally (e.g., bipolar disorder, schizophrenia, and the personality disorders) not only have contributed to a deeper understanding of the dysfunctions but also have educated the public about the causes and consequences of mental disorder (see Cicchetti & Cannon, 1999; Cicchetti & Crick, 2009a, 2009b; Miklowitz & Cicchetti, 2006, 2010; Tackett & Sharp, 2014). Advances in genomics, GxE interactions, and epigenetics; growth in our understanding of neurobiology, neural plasticity, and resilience; and progress in the development of methodological and technological tools, including brain imaging, neural circuitry, hormone assays, immunology, social and environmental influences on brain development, and statistical analysis of developmental change, pave the way for interdisciplinary and for multiple levels of analysis research programs that will significantly increase the knowledge base of the development and course of maladaptation, psychopathology, and resilience. Moreover, randomized control prevention and intervention trials are being conducted based on theoretical models and efforts to elucidate the mechanisms and processes contributing to developmental change at both the biological and psychological levels (Belsky & van IJzendoorn, 2015; Cicchetti & Gunnar, 2008). Despite the significant advances that have occurred in the field of developmental psychopathology, much important work lies ahead. Undoubtedly these future developments will build on the venerable contributions of the past; however, as work in the field becomes increasingly interdisciplinary, multilevel, and technologically sophisticated, it is essential that even more emphasis be directed toward the process of development (Harter, 2006; Sroufe, 2007, 2013). It is not only genes and environments but also the cumulative developmental history of the individual that influences how future development will unfold (Sroufe, 2007, 2013). Developmental psychopathologists have incorporated concepts and methods derived from other disciplinary endeavors that are too often isolated from each other, thereby generating advances in knowledge that might have been missed in the absence of cross-disciplinary dialogue. The continuation and elaboration of the mutually enriching interchanges that have occurred within and across disciplines interested in normal and abnormal development not only will enhance the science of developmental psychopathology but also will increase the benefits to be derived for individuals with high-risk conditions or mental disorders, families, and society as a whole. Dante Cicchetti, Ph.D. Minneapolis, MN January 2015

References Belsky, J., & van IJzendoorn, M. (2015). What works for whom? Genetic moderation of intervention efficacy. [Special Section]. Development and Psychopathology, 27, 1–6.

Cicchetti, D. (1984). The emergence of developmental psychopathology. Child Development, 55(1), 1–7. Cicchetti, D. (1990). A historical perspective on the discipline of developmental psychopathology. In J. Rolf, A. Masten, D. Cicchetti, K. Nuechterlein, & S. Weintraub (Eds.), Risk and protective factors in the development of psychopathology (pp. 2–28). New York, NY: Cambridge University Press. Cicchetti, D. (1993). Developmental psychopathology: Reactions, reflections, projections. Developmental Review, 13, 471–502. Cicchetti, D., & Cannon, T. (1999). Neurodevelopmental processes in the ontogenesis and epigenesis of psychopathology. Development and Psychopathology, 11, 375–393. Cicchetti, D., & Cohen, D. (Eds.). (1995). Developmental psychopathology (Vols. 1–2). New York, NY: Wiley. Cicchetti, D., & Cohen, D. (Eds.). (2006). Developmental psychopathology (2nd ed., Vols. 1– 3). New York, NY: Wiley. Cicchetti, D., & Crick, N. R. (Eds.) (2009a). Precursors of and diverse pathways to personality disorder in children and adolescents. [Special Issue, Part 1]. Development and Psychopathology, 21(3), 683–1030. Cicchetti, D., & Crick, N. R. (Eds.). (2009b). Precursors of and diverse pathways to personality disorder in children and adolescents. [Special Issue, Part 2]. Development and Psychopathology, 21(4), 1031–1381. Cicchetti, D., & Gunnar, M. R. (2008). Integrating biological processes into the design and evaluation of preventive interventions. Development and Psychopathology, 20, 737–743. Cicchetti, D., & Gunnar, M. R. (Eds.). (2009). Meeting the challenge of translational research in child psychology: Minnesota symposia on child psychology (Vol. 35). New York, NY: Wiley. Cicchetti, D., & Sroufe, L. A. (2000). The past as prologue to the future: The times they've been a changin'. Development and Psychopathology, 12, 255–264. Cicchetti, D., & Toth, S. L. (1991). The making of a developmental psychopathologist. In J. Cantor, C. Spiker, & L. Lipsitt (Eds.), Child behavior and development: Training for diversity (pp. 34–72). Norwood, NJ: Ablex. Cicchetti, D., & Toth, S. L. (Eds.). (2006). Translational research in developmental psychopathology. [Special Issue]. Development and Psychopathology, 18(3), 619–933. Cicchetti, D., & Toth, S. L. (2009). The past achievements and future promises of developmental psychopathology: The coming of age of a discipline. Journal of Child Psychology and Psychiatry, 50, 16–25.

Harter, S. (2006). Self-processes and developmental psychopathology. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology (2nd ed., 370–418). New York, NY: Wiley. Masten, A. S. (2014). Ordinary magic: Resilience in development. New York, NY: Guilford Publications, Inc. Miklowitz, D. J., & Cicchetti, D. (2006). Toward a life span developmental psychopathology perspective on bipolar disorder. Development and Psychopathology, 18, 935–938. Miklowitz, D. J., & Cicchetti, D. (Eds.). (2010). Bipolar disorder: A developmental psychopathology approach. New York, NY: Guilford. Rutter, M. (1986). Child psychiatry: The interface between clinical and developmental research. Psychological Medicine, 16, 151–169. Rutter, M. (1987). Psychosocial resilience and protective mechanisms. American Journal of Orthopsychiatry, 57, 316–331. Rutter, M. (2012). Resilience as a dynamic concept. Development and Psychopathology, 24, 335–344. Rutter, M., & Garmezy, N. (1983). Developmental psychopathology. In E. M. Hetherington (Ed.), Handbook of child psychology (pp. 774–911). New York, NY: Wiley. Rutter, M., & Sroufe, L. A. (2000). Developmental psychopathology: Concepts and challenges. Development and Psychopathology, 12, 265–296. Sroufe, L. A. (1990). Considering normal and abnormal together: The essence of developmental psychopathology. Development and Psychopathology, 2, 335–347. Sroufe, L. A. (2007). The place of development in developmental psychopathology. In A. Masten (Ed.), Multilevel dynamics in developmental psychopathology pathways to the future: The Minnesota symposia on child psychology (pp. 285–299). Mahwah, NJ: Erlbaum. Sroufe, L. A. (2013). The promise of developmental psychopathology. Development and Psychopathology, 25, 1215–1224. Sroufe, L. A., & Rutter, M. (1984). The domain of developmental psychopathology. Child Development, 55, 17–29. Tackett, J. L., & Sharp, C. (2014). A developmental psychopathology perspective on personality disorder. [Special Issue]. Journal of Personality Disorders, 28, 1–179. Weiss, P. (1969). Principles of development. New York, NY: Hafner.

Contributors Manuel Barrera, Jr., PhD Arizona State University Tempe, Arizona Marc H. Bornstein, PhD National Institute of Child Health and Human Development Bethesda, Maryland Keith B. Burt, PhD University of Vermont Burlington, Vermont Jocelyn S. Carter, PhD DePaul University Chicago, Illinois Juan F. Casas, PhD University of Nebraska Omaha, Nebraska Edith Chen, PhD Northwestern University Evanston, Illinois Xinyin Chen, PhD University of Pennsylvania Philadelphia, Pennsylvania Dante Cicchetti, PhD Institute of Child Development University of Minnesota Minneapolis, Minnesota J. Douglas Coatsworth, PhD Colorado State University

Fort Collins, Colorado Nicki R. Crick, PhD University of Minnesota Minneapolis, Minnesota Jessica Dollar, PhD University of North Carolina Greensboro, North Carolina Sophia Duffy, PhD DePaul University Chicago, Illinois Katherine B. Ehrlich, PhD Northwestern University Evanston, Illinois Gary W. Evans, PhD Cornell University Ithaca, New York Nathan A. Fox, PhD University of Maryland College Park, Maryland Nancy A. Gonzales, PhD Arizona State University Tempe, Arizona Rebecca Goodvin, PhD Western Washington University Bellingham, Washington Kathryn Grant, PhD DePaul University Chicago, Illinois Julie A. Gravener-Davis, PhD Mt. Hope Family Center

Rochester, New York Elizabeth D. Handley, PhD Mt. Hope Family Center Rochester, New York Stephen P. Hinshaw, PhD University of California Berkeley, California Ernest N. Jouriles, PhD Southern Methodist University Dallas, Texas Robert-Paul Juster, MSc, PhD McGill University Montreal, Canada Chrystyna D. Kouros, PhD Southern Methodist UniversityDallas, TX Dave Lanoix, PhD Centre for Studies on Human Stress Montreal, Canada Anna S. Lau, PhD University of California Los Angeles, California Cindy H. Liu, PhD Harvard Medical School Boston, Massachusetts Sonia J. Lupien, PhD University of Montreal Montreal, Canada Ian Mahar, BSc McGill University Montreal, Canada

Andres G. Martinez, PhD University of California Berkeley, California Ann S. Masten, PhD Institute of Child Development University of Minnesota Minneapolis, Minnesota Renee McDonald, PhD Southern Methodist University Dallas, Texas Bruce S. McEwen, PhD Rockefeller University New York, New York Naguib Mechawar, PhD McGill University Montreal, Canada Gregory E. Miller, PhD Northwestern University Evanston, Illinois Dianna Murray-Close, PhD University of Vermont Burlington, Vermont Velma M. Murry, PhD Vanderbilt University Nashville, Tennessee Charles A. Nelson, PhD Harvard University Boston, Massachusetts and Boston Children's Hospital

Boston, Massachusetts David A. Nelson, PhD Brigham Young University Provo, Utah Jamie M. Ostrov, PhD University at Buffalo, The State University of New York Buffalo, New York Isabelle Ouellet-Morin, PhD University of Montreal Montreal, Canada Christie L. M. Petrenko, PhD Mt. Hope Family Center Rochester, New York Robert C. Pianta, PhD University of Virginia Charlottesville, Virginia Martin Picard, PhD Columbia University New York, New York Armando A. Pina, PhD Arizona State University Tempe, Arizona Pierrich Plusquellec, PhD University of Montreal Montreal, Canada Robert M. Post, MD Bipolar Collaborative Network Bethesda, Maryland Zoltan Sarnyai, PhD, MD James Cook University

Townsville, Australia Teresa Seeman, PhD University of California Los Angeles, California Shireen Sindi, PhD McGill University Montreal, Canada Ellen A. Skinner, PhD Portland State University Portland, Oregon Nathan Grant Smith, PhD University of Houston Houston, Texas Juliana Souza-Talarico, PhD University of São Paulo São Paulo, Brazil Margaret Beale Spencer, PhD University of Chicago Chicago, Illinois Cynthia Stifter, PhD Pennsylvania State University University Park, Pennsylvania Dena Phillips Swanson, PhD University of Rochester Rochester, New York Ross A. Thompson, PhD University of California Davis, California Patrick H. Tolan, PhD University of Virginia

Charlottesville, Virginia Sheree L. Toth, PhD Mt. Hope Family Center Rochester, New York Martha E. Wadsworth, PhD Pennsylvania State University University Park, Pennsylvania Charles H. Zeanah, PhD Tulane University New Orleans, Louisiana Melanie J. Zimmer-Gembeck, PhD Griffith University Gold Coast, Australia

Chapter 1 Childhood Adversity and Adult Physical Health Katherine B. Ehrlich, Gregory E. Miller, and Edith Chen DEFINING CHILDHOOD ADVERSITY Child Maltreatment Socioeconomic Disadvantage Summary DEFINING HEALTH OUTCOMES CHILDHOOD ADVERSITY AND LATER DISEASE: EPIDEMIOLOGICAL EVIDENCE Maltreatment and Later Disease Socioeconomic Disadvantage and Later Disease Other Forms of Adversity and Later Disease Limitations and Alternative Explanations CONCEPTUAL MODELS LINKING CHILDHOOD ADVERSITY TO ADULT PHYSICAL HEALTH Cumulative Models Diathesis–Stress Models Differential Susceptibility Buffering Models Developmental Trajectory Models Developmental Cascades BIOLOGICAL INTERMEDIARIES LINKING EARLY ADVERSITY TO ADULT PHYSICAL HEALTH Hypothalamic-Pituitary-Adrenocortical Axis Allostatic Load Telomeres Epigenetics Evidence for the Role of Epigenetic Modifications as a Biological Intermediary Epigenetic Modifications as a Biological Intermediary: Considerations and Caveats

Inflammation CONCLUDING COMMENTS AND DIRECTIONS FOR FUTURE RESEARCH Expanding Research to Other Periods of Development Strengthening Study Designs Buffers and Protective Factors Translational Implications Linking Research on Developmental Psychopathology and Health Psychology Conclusions REFERENCES A consensus in the developmental psychopathology literature is that the experience of early adversity—particularly the experience of chronic, uncontrollable stress—is a risk factor for a diverse set of poor outcomes across development (e.g., Appleyard, Egeland, van Dulmen, & Sroufe, 2005; Poulton et al., 2002; Shonkoff et al., 2012). In the last three decades, a growing number of studies have provided convincing evidence to conclude that adversity in childhood has a lasting influence on adult physical health, particularly chronic diseases associated with aging, like cardiovascular disease, diabetes, arthritis, and some cancers (Gluckman & Hanson, 2006; Miller, Chen, & Parker, 2011; Repetti, Taylor, & Seeman, 2002). This susceptibility to the chronic diseases of aging resulting from early adversity has been identified in diverse samples with a range of adverse risk factors, including socioeconomic disadvantage, maltreatment, and chaotic family environments. These findings suggest that these early stressful experiences leave a “biological residue,” manifesting in physical health problems in adulthood. This chapter provides a review of the current knowledge linking childhood adversity to adult physical health. We first provide a discussion of childhood adversity and the varied ways that children encounter stressful experiences in their daily lives. Then we review evidence for links between childhood adversity and chronic diseases of aging, with a focus on cardiovascular disease and metabolic disorders, where most of the research to date has occurred. We then describe conceptual models that help guide empirical investigation of the processes through which early adversity influences later health. We also review biological mechanisms that might play an intermediary role in translating psychosocial stress into health problems, which can help explain how early adversity “gets under the skin” and influences the onset of disease in adulthood. Our goal is to highlight the advantages and limitations in the conclusions we can draw from research on these biological processes. Finally, we end this chapter with a series of important research questions that should be a focus in the next generation of empirical investigation on early adversity and adult physical health.

Defining Childhood Adversity

There are a number of ways that children could experience significant adversity. In this chapter, we focus on adversity that is both chronic and severe in nature. We define chronic adversity as one that remains present in the child's life over a significant period of time (e.g., lack of material resources due to poverty). Adversity can also be chronic when children experience lingering threat over the possibility of a repeated stressful experience (e.g., stress resulting from a traumatic experience, such as abuse, that could reoccur) or aftereffects of adversities that create severe disruptions to daily life (e.g., long-term displacement resulting from a natural disaster). Adversity is considered severe when it results in a profound unsettling of normative childhood experiences and threatens the well-being of the child (e.g., exposure to gang violence while living in poverty). Studies that have attempted to estimate the prevalence of adverse life experiences in childhood have found that these stressors actually are fairly common. Kessler and colleagues (Kessler, Davis, & Kendler, 1997) categorized early-life adversities into four domains, including (a) interpersonal traumas (e.g., rape, physical attacks); (b) loss events (e.g., death of a parent); (c) parental psychopathology (e.g., parental depression, antisocial personality disorder); and (d) a miscellaneous category of stressful life events (e.g., life-threatening accidents, man-made disasters). Using the National Comorbidity Survey, a nationally representative sample of families in the United States, Kessler et al. found that by the age of 16, nearly 75% of children have experienced at least one significant adversity and approximately 50% of children have experienced multiple adversities. In this sample, the most frequent adversities that children faced were maternal depression, paternal verbal abuse, paternal substance abuse, and parental separation or divorce. Further, a startling number of children currently face adversity in impoverished or economically stressed conditions. According to the Children's Defense Fund (2012), 22% of children in the United States were living in poverty in 2010, meaning that a family of four earns less than $22,350 a year (U.S. Department of Health and Human Services, 2011). An additional 22% of children live in low-income families, defined as less than 200% of the federal poverty level (or less than $44,700 for a family of four). Children in low-income families may be exposed to many of the same stressors that children living below the poverty line face, including dangerous neighborhoods, material deprivation, parental underemployment, and familial mental health problems, all contributing to strained family relationships. Overwhelmingly, research on the links between early adversity and later physical health has focused on two types of adversity: child maltreatment and socioeconomic disadvantage. Although these experiences of early adversity fall under the definition of adversity proposed above, they differ from each other in some critical aspects, including the nature and source of threatening experiences; the duration, frequency, and severity of those experiences; and the opportunities for coping. Yet maltreatment and socioeconomic disadvantage share several overlapping qualities, which may include cold, insensitive parenting; harsh discipline; exposure to conflict and violence; limited access to resources; and uncertainty of future environmental stability (Repetti et al., 2002). Next we take a closer look at several forms of adversity that are thought to be particularly detrimental for later physical health.

Child Maltreatment A fundamental component of attachment theory is the notion that individuals develop representations (or internal working models) that reflect the extent to which a caregiver serves as a secure base for exploration and as a safe haven when needed for comfort and support (Bowlby, 1969/1982, 1973). These representations are experience-based, and they are developed over the first year of life in response to repeated interactions with a caregiver. When these caregiving experiences include neglect or abuse, children learn that their caregiver is not an available secure base or safe haven. Further, these children may develop unusual behaviors, characterized by odd, fearful, and disorganized patterns of interactions with a caregiver—characteristics that emerge when the caregiver becomes their primary source of fear and support (Lyons-Ruth & Jacobvitz, 2008). These behavioral responses reflect the child's inability to cope with a chaotic environment. Importantly, children who are maltreated by caregivers grow up without the experience of knowing that a reliable caregiver is available when needed—an aspect of the family emotional climate that plays a fundamental role in children's abilities to regulate distress and cope with negative emotions (Bowlby, 1988). Child maltreatment is a serious public health concern that poses significant mental and physical health burdens on its victims (Cicchetti & Toth, 2005) as well as substantial economic burdens on society as a whole (Fang, Brown, Florence, & Mercy, 2012). Estimates suggest that almost 700,000 children were victimized in the United States in 2011, almost half of whom were under 6 years old at the time of the abuse (U.S. Dept. of Health and Human Services, 2011). Of course, because a large number of cases are unreported and a third of reported cases are not investigated (Cicchetti & Toth, 2005), the actual rate of child maltreatment may in fact be higher. The vast majority of cases (78%) included children suffering from neglect, but each year, hundreds of thousands of children are also victims of physical, sexual, or psychological abuse. Alarmingly, childhood maltreatment is perpetrated most often by primary caregivers, with over 80% of cases involving abuse by one or more parents. Explanations for the causes of child maltreatment widely recognize that the phenomenon is multiply determined. Attempts to characterize the contexts of maltreatment have cited parental factors (e.g., mental illness, substance abuse), child factors (e.g., difficult temperament, disruptive behavior), interactional factors—that is, the dynamic transactions that take place between parents and children that might incite abuse (e.g., a child's aggressive behavior elicits physical punishment from parents that subsequently escalates to abuse), and environmental factors (e.g., cultural attitudes, poverty; Belsky, 1993; Cicchetti & Toth, 2005).

Socioeconomic Disadvantage Children growing up with socioeconomic disadvantage have a higher probability of being exposed to stressful conditions across virtually every domain of daily life. Many of these stressful conditions center on the lack of security for basic resources, such as food and housing. For example, over 10% of children live in “food-insecure” families who struggle to provide enough food to meet children's daily nutritional needs, and over 20 million children receive free or reduced-cost lunches at school—meals they are eligible to receive because of

their family's scarce financial resources (Children's Defense Fund, 2012). Children in poverty often do not have adequate access to medical care, resulting in infrequent visits to the doctor and risk for serious complications from untreated illnesses. Children growing up in poverty are at risk for low educational attainment, meager job success, and incarceration (Duncan, Kalil, & Ziol-Guest, 2008). Notably, children in poverty also face neighborhood stressors, and they are at increased risk for becoming victims of violence, theft, and other crimes (Ross & Mirowsky, 2001). Further, their caregivers are burdened by numerous demands, such as multiple part-time jobs, unaccommodating schedules, and psychosocial stress brought on by their lack of resources, making it more difficult for them to serve as sensitive sources of support for their children. The accumulation of deficits in resources—both material and psychological—cluster to create an environment that is ill-equipped to foster children's healthy development. Children living in disadvantaged families can face daily threats to their basic needs and be exposed to additional stressors that can jeopardize their mental and physical well-being. Presumably, other adverse experiences in childhood (including many of the factors that Kessler et al. 1997 identified) have repercussions for adult health. To date, however, these links between exposure to other early adversities besides poverty and maltreatment and adult health have not been a primary focus in the literature. For example, with the exception of a handful of studies, we know very little about how the loss of a parent, the experience of severe interparental conflict, or exposure to neighborhood stress plays a role in shaping chronic disease outcomes in adulthood. Children are exposed to a wide array of adverse experiences, and it will be important to understand the extent to which each of these adversities is associated with long-term physical health outcomes.

Summary Chronic adversity is a problem for many children in the United States. So far, the literature on health consequences has focused mainly on maltreatment and disadvantage. But an important question for future studies to consider is the impact of other significant adversities (e.g., interparental conflict, long-term separations from parents, natural disasters) on adult physical health. One challenge to this line of research, however, is that many forms of adversity—and socioeconomic disadvantage and maltreatment in particular—co-occur more often than would be expected by chance alone (Crouch, Hanson, Saunders, Kilpatrick, & Resnick, 2000), thus creating a unique challenge to parsing out the effects of different forms of adversity on later physical health outcomes. The use of a variety of statistical techniques (e.g., variable-centered and person-centered approaches to data analysis) as well as comprehensive models (described in more detail later in this chapter) offer opportunities to circumvent the challenges associated with identifying the complex connections between early adversity and adult physical health.

Defining Health Outcomes Studies documenting the link between early adversity and adult physical health have assessed everything from the frequency of minor physical symptoms like headaches and constipation to

rates of disorders and death from various conditions. As much as possible we focus on disease morbidity (the diagnosis of a disease or a clinical manifestation of it) and mortality rather than symptom reports. The advantages of focusing on disease endpoints and death are that they can be ascertained objectively and are viewed as reflecting differences in an underlying disease process. Further, these outcomes are much simpler to interpret than symptoms reported by patients, which tend to be heavily shaped by individual differences in symptom perception, labeling, and reporting (Feldman, Cohen, Doyle, Skoner, & Gwaltney, 1999). Our focus in this chapter rests on the major chronic diseases associated with aging, including studies of risk for cardiovascular disease, stroke, cancer, diabetes, and metabolic syndrome, which is a precursor of cardiovascular disease and diabetes (Grundy et al., 2005). These diseases account for the vast majority of suffering and disability in the United States (and well over 1 million deaths in the United States each year; Murphy, Xu, & Kochanek, 2012). Further, the economic burden associated with health care costs for treating these chronic diseases— which amounts to more than three-quarters of the nation's health care spending—accounts for approximately $300 billion annually (DeVol & Bedroussian, 2007).

Childhood Adversity and Later Disease: Epidemiological Evidence Mounting evidence suggests that early childhood adversity is associated with a number of chronic health problems later in life, including cardiovascular disease, diabetes, cerebrovascular disease, and even some cancers (Adler & Rehkopf, 2008; Gluckman & Hanson, 2006; Shonkoff et al., 2012). This quickly growing body of research has implicated psychosocial stress as a primary mediator of adversity's association with morbidity and mortality from chronic diseases of aging (Matthews & Gallo, 2011; Shonkoff, Boyce, & McEwen, 2009). In the sections that follow, we review epidemiological evidence for the role of early adversity in contributing to later disease. We focus mainly on the role of maltreatment and socioeconomic disadvantage, two forms of adversity that have been studied extensively in relation to later health problems. Where applicable, we review available literature documenting other forms of early adversity that have been linked to chronic diseases of aging. When possible, we selected studies that used nationally representative studies with large sample sizes, assessment of possible confounds, prospective study designs, and objective measures of health outcomes in adulthood. In some domains, these criteria could not be met within any single study, so we review available evidence and note the methodological concerns. Further, our discussion is restricted to studies that link childhood adversity to adult physical health. Although a growing body of research has documented links between exposure to stress in childhood and childhood disease outcomes, these studies are beyond the scope of this chapter.

Maltreatment and Later Disease Studies of the long-lasting sequelae resulting from maltreatment in childhood have identified

robust links between childhood abuse and chronic diseases of aging, including cardiovascular disease and diabetes. Evidence for these links comes from studies that focus on different forms of maltreatment, including physical abuse, sexual abuse, and neglect (e.g., Anda et al., 2006; Goodwin & Stein, 2004). The vast majority of these studies used retrospective reports about maltreatment, which presents some methodological concerns about the accuracy of the reports. Next we review a selection of studies that have examined links between maltreatment and specific disease outcomes. Maltreatment and Cardiovascular Disease Links between early maltreatment and cardiovascular disease are well documented (e.g., Batten, Aslan, Maciejewski, & Mazure, 2004; Dong et al., 2004; Draper et al., 2008; for metaanalytic findings, see Irish, Kobayashi, & Delahanty, 2010; Wegman & Stetler, 2009). For example, using data from the National Comorbidity Survey, Goodwin and Stein (2004) found that recalled sexual abuse in childhood was associated with increased risk for cardiovascular disease, including heart attacks. Similarly, findings from the Nurses' Health Study revealed links between women's recollections of childhood physical and sexual abuse and cardiovascular disease (Rich-Edwards et al., 2012). In this study, severe physical abuse was associated with a 46% greater risk of having a cardiovascular event (e.g., heart attack, stroke). Similarly, sexual abuse was associated with 56% increased odds of having a cardiovascular event. Links between maltreatment and cardiovascular disease have been found even in complex analyses that include possible alternative explanations, including adult health risk behaviors (e.g., smoking), stressors in adulthood (e.g., daily stress, low educational attainment), depression, and childhood stressors other than maltreatment (e.g., parental divorce). Even after accounting for these potential confounding variables, Fuller-Thomson, Brennenstuhl, and Frank (2010) found that recalled childhood physical abuse was associated with 45% greater likelihood of being diagnosed with heart disease. Felitti and colleagues (1998) have suggested that some of these confounding variables may serve as mediating mechanisms in the progression of heart disease (e.g., the experience of maltreatment may lead individuals to engage in unhealthy lifestyles), but the findings from many of these studies indicate that maltreatment still accounts for some of the risk for cardiovascular disease even after accounting for health behaviors, which may point to other mechanisms linking early maltreatment to cardiovascular disease. Several investigations have identified gender-specific links between maltreatment and cardiovascular disease (e.g., Batten et al., 2004; Draper et al., 2008; Fuller-Thomson, Bejan, Hunter, Grundland, & Brennenstuhl, 2012; Goodwin & Stein, 2004). For example, FullerThomson et al. (2012) found that childhood sexual abuse was associated with a greater risk for heart attacks for men but not for women. In contrast, in the National Comorbidity Survey, Goodwin and Stein (2004) and Batten et al. (2004) found connections between maltreatment and cardiovascular disease for women but not for men. Unfortunately, insufficient evidence exists currently to make broad conclusions about gender-specific risk for cardiovascular disease resulting from early maltreatment. Nevertheless, it will be important for future

research to examine whether the link between maltreatment and cardiovascular disease differs for men and women. It may be that certain forms of maltreatment (e.g., sexual abuse or neglect) are associated with unique risk outcomes for men versus women. These findings suggest that exposure to early maltreatment is linked to adult cardiovascular disease, but the studies' methodological limitations make it difficult to conclude that maltreatment causally contributes to heart disease. For example, these studies rely exclusively on retrospective self-reports of maltreatment (often with a single question about whether participants were ever abused or neglected), which calls into question the accuracy of the assessment due to possible memory errors or informant biases. In addition, most of these studies use participant reports about their medical diagnoses, which are less trustworthy than medical records. These study weaknesses are not easily discounted, and they limit our ability to make strong conclusions about the role of maltreatment in contributing to disease outcomes. To address these concerns, future studies will need to identify and control for factors that could give rise to spurious associations between maltreatment and subsequent cardiovascular disease (e.g., neuroticism). The use of strong methodological designs, including studies from administrative databases that can utilize objectively verified records of maltreatment and cardiovascular disease, would help assuage skepticism about the link between child maltreatment and adult cardiovascular disease. Maltreatment and Metabolic Risk Childhood maltreatment has been found to be predictive of diabetes and other metabolic abnormalities in adulthood (e.g., Danese et al., 2009; Felitti et al., 1998; Goodwin & Stein, 2004; Rich-Edwards et al., 2010; Thomas, Hyppönen, & Power, 2008). In recent investigations, researchers have examined a cluster of metabolic abnormalities known as the metabolic syndrome (Cornier et al., 2008; Grundy et al., 2005; previously known as Syndrome X). Metabolic syndrome is increasingly viewed as a precursor to cardiovascular disease and diabetes and reflects sedentary lifestyles and overnutrition. Its connection to potentially fatal cardiac events (e.g., heart attacks, strokes) and high prevalence in modern society (estimates suggest that around 30% of adults in the United States meet criteria for metabolic syndrome; Cornier et al., 2008) make it a promising predisease marker for individuals who have not yet developed chronic diseases. In the first study of links between early maltreatment and diabetes, Felitti et al. (1998) found that diabetes was more prevalent when individuals reported four or more indicators of childhood risk (including maltreatment and family dysfunction). Similarly, in a community sample of women in New Zealand, Romans, Belaise, Martin, Morris, and Raffi (2002) found that childhood abuse was associated with increased diabetes prevalence. Further evidence for a link between maltreatment and metabolic risk comes from the National Comorbidity Survey (Goodwin & Stein, 2004). In this sample, recalled childhood neglect (but not physical or sexual abuse) was associated with increased odds of having diabetes. Despite variations in definitions of maltreatment (e.g., neglect versus abuse), these studies suggest that the experience of poor caregiving conditions in childhood is a risk factor for diabetes in adulthood. Interestingly, Goodwin and Stein's (2004) findings of the specific type of

maltreatment experience that was associated with metabolic risk suggests that certain types of maltreatment may be particularly influential in shaping diabetes onset. To date, only a handful of studies have examined whether maltreatment in childhood is a predictor of prediabetes metabolic risk in adulthood, but preliminary evidence provides some support for this link. Midei, Matthews, Chang, and Bromberger (2013) examined connections among emotional, physical, and sexual abuse and the onset of metabolic syndrome in a sample of women in middle adulthood. In this study, childhood physical abuse, but not sexual or emotional abuse, was associated with the onset of metabolic syndrome over a 7-year span. In one of the only prospective studies of maltreatment and metabolic risk, Danese et al. (2009) examined the separate predictive roles of child maltreatment, social isolation, and socioeconomic status (SES) as well as the cumulative exposure to adversity across these domains. Using the Dunedin Multidisciplinary Health and Development Study, Danese et al. grouped participants into three categories according to their exposure to maltreatment in childhood: (a) no maltreatment; (b) probable maltreatment—meaning that children experienced one out of five indicators of maltreatment; and (c) definite maltreatment, wherein children experienced two or more indicators of maltreatment. Although low SES and social isolation were independently predictive of metabolic risk, child maltreatment was not reliably associated with higher rates of metabolic risk factors. Somewhat surprisingly, adults who were categorized as “probable maltreatment” had elevated risk for metabolic abnormalities, but adults who were categorized as “definite maltreatment” were not at increased risk. This finding is particularly noteworthy, as it is the only study to our knowledge that has multiple informant ratings of maltreatment assessed when participants were children and does not rely on retrospective reports. Given that participants were still relatively young (32 years old) and healthy at the time of assessment of metabolic indicators, it is not entirely surprising that maltreatment was not a reliable risk factor for metabolic disruptions. It may be the case that maltreatment takes longer to manifest itself in metabolic disruptions that will be seen when these participants are older. Yet it is also possible that other studies that have used retrospective self-reports of maltreatment have overstated the link between early maltreatment and adult metabolic risk. Thus, although it appears that early maltreatment is a risk factor for both diabetes and prediabetes metabolic risk, additional research with multiple informants and prospective study designs will be important for shedding light on the extent to which harsh caregiving experiences and maltreatment are predictive of later metabolic disruptions. Maltreatment and Cancer Compared to the well-established literature that examines early maltreatment and cardiovascular disease and diabetes, much less is known about the extent to which maltreatment in childhood confers additional risk for cancer. A handful of studies have studied this link using retrospective study designs, with mixed findings (e.g., Brown et al., 2010; Draper et al., 2008; Felitti et al., 1998; Fuller-Thomson & Brennenstuhl, 2009; Morton, Schafer, & Ferraro, 2012). In the first reported examination of this link, Felitti et al. (1998) found that individuals were at increased risk for cancer when they experienced four or more adverse childhood risk factors. Similarly, in the National Survey of Midlife Development in

the United States (MIDUS), Morton and colleagues (2012) examined links between childhood abuse and cancer and found that emotional and physical abuse were associated with increased risk of cancer at midlife. These links remained significant even after controlling for potential confounding variables, including age, race, SES, and health-related behaviors (e.g., smoking). Additional evidence for a link between childhood maltreatment and cancer comes from a sample of over 13,000 Canadians who took part in the Canadian Community Health Survey (Fuller-Thomson & Brennenstuhl, 2009). In this sample, recalled physical abuse was associated with 47% higher odds of cancer, even when adjusting for risk factors, such as childhood stressors, adult health behaviors, and adult SES. Morton et al. and Fuller-Thomson and Brennenstuhl (2009) relied on studies that used self-reports of abuse and cancer, which raises interpretational ambiguities, as we noted earlier. However, self-reports about cancer diagnosis may be less subject to false reports than other disease outcomes or symptom reports. Prospective studies and studies utilizing administrative databases with objective measures of maltreatment and disease diagnoses will bolster support for the notion that child maltreatment is a risk factor for adult cancer diagnosis. Other studies, however, have not found links between early maltreatment and later cancer risk. In a sample of over 21,000 older adults, Draper et al. (2008) examined links between childhood physical and sexual abuse and adult mental and physical health outcomes. In this study, although adults who had been abused reported worse mental and physical health, abused participants were not at greater risk for experiencing cancer in adulthood. Similarly, Korpimäki, Sumanen, Silanmäki, and Mattila (2010) did not find a link between maltreatment and cancer in a large epidemiological study in Finland. A number of factors could account for these inconsistent findings. For one, researchers vary in their definitions of maltreatment: Some studies include a range of harsh parenting indices, whereas other studies take a more explicit approach by asking participants if they had been abused as children. Moreover, there is considerable variability across studies in sample characteristics and analytical approaches (e.g., the extent to which possible confounding variables are included in statistical models, whether maltreatment is considered as a continuous or binary variable). Another difficulty is that cancer prevalence rates are lower than the rates of heart disease and diabetes, thus making it a more difficult outcome to predict. Further, cancer is a much more heterogeneous disease compared to diabetes and cardiovascular disease; this heterogeneity may explain some of the inconsistent findings, particularly if studies vary in their focus on particular cancers versus all cancer diagnoses. Additional investigation into the connections between maltreatment and cancer risk will provide insight into whether the experience of maltreatment in childhood is a risk factor for later cancer diagnosis. Summary of Research on Early Maltreatment and Later Disease Consistent with the notion that early exposure to adversity is linked to many of the chronic diseases of aging, preliminary evidence suggests that maltreatment in childhood may be a risk factor for poor health outcomes in adulthood. In the vast majority of these studies, however, researchers have relied on retrospective reports of maltreatment, which are subject to significant reporting biases. Moreover, these reporting biases might not be random: It is

possible that individuals with poor health remember their prior experiences in negatively biased ways. Further, many of these studies use self-reports of physician diagnoses of diseases. Although we prefer these markers of physical health over informants' own symptom reports, these assessments are nevertheless vulnerable to inaccurate reports. Additional research will be useful for shedding light on the specific contexts in which maltreatment is likely to be a serious risk for later health problems. For example, it is possible that certain forms of maltreatment are more likely to contribute to health problems for women but not for men, or for people from particular racial or ethnic backgrounds. Another lingering question concerns the relative weight of various forms of maltreatment in the progression of long-term health problems: Do all forms of maltreatment pose risk for health problems, or are certain forms of abuse or neglect especially likely to have a negative effect on later physical health? Moreover, the experience of a particular type of maltreatment (e.g., physical abuse) and its impact on later health may depend, in part, on the cultural context in which it occurs. For example, in cultures that condone corporal punishment, links between childhood physical abuse and later health may be less identifiable. These research questions await empirical examination.

Socioeconomic Disadvantage and Later Disease The literature on the role of childhood socioeconomic disadvantage in shaping risk for later disease and mortality risk is extensive and suggests that, like early maltreatment, it is a risk for poor health outcomes in adulthood. In particular, low SES in childhood is a risk factor for early mortality, cardiovascular disease, diabetes, and some cancers (Cohen, Janicki-Deverts, Chen, & Matthews, 2010; Galobardes, Lynch, & Smith, 2004, 2008; Kumari, Head, & Marmot, 2004). When studying the long-term health outcomes associated with childhood socioeconomic disadvantage, an important factor to consider is that SES tends to be stable across the lifespan (Hertzman, 1999). Such stability makes attempts to disentangle the effects of early versus adult SES on health particularly difficult. Given that adult SES is a robust predictor of morbidity and mortality from chronic diseases of aging (Adler & Rehkopf, 2008; Lynch & Smith, 2005), many studies of early childhood SES statistically control for adult SES in analyses, thus allowing for better insight into the role that early-life experiences play in influencing the risk for later disease. In order to avoid multicollinearity issues associated with highly stable childhood and adult SES estimates, recently researchers have developed statistical modeling techniques to better account for early and adult SES (e.g., Nandi, Glymour, Kawachi, & VanderWeele, 2012). Socioeconomic Disadvantage and Cardiovascular Disease In their systematic review of the literature on SES and cardiovascular disease, Galobardes, Smith, and Lynch (2006) concluded that low SES in childhood is a significant risk for cardiovascular disease in adulthood. Even when adult SES was controlled in statistical analyses, the link between childhood SES and disease risk was still significant, accounting for

20% to 40% of the risk for cardiovascular disease. In a large-scale investigation of socioeconomic disadvantage and cardiovascular disease, Claussen, Smith, and Thelle (2003) found that cardiovascular disease-related mortality was associated with childhood SES. Similarly, researchers found evidence for the role of early socioeconomic disadvantage in predicting cardiovascular disease in the British Whitehall II study (Singh-Manoux, Ferrie, Chandola, & Marmot, 2004). Another study tracked the onset of coronary heart disease over a period of 40 years in a sample of physicians who graduated from medical school at Johns Hopkins University (Kittleson et al., 2006). Even among these educated, affluent physicians, childhood adversity was associated with worse health in adulthood. The rates of cardiovascular disease by age 50 were 2.4 times higher in physicians who were raised in households that were low versus high in SES. These findings offer dramatic evidence in support of a link between socioeconomic disadvantage and cardiovascular disease. Socioeconomic Disadvantage and Metabolic Risk A number of studies have examined the role of early socioeconomic disadvantage and risk for metabolic abnormalities, including the metabolic syndrome and diabetes (e.g., Agardh et al., 2007; Brunner et al., 1997; Kumari et al., 2004; Parker et al., 2003). Support for this link has been somewhat mixed, although the majority of studies have found at least some evidence of a link between early low socioeconomic standing and greater risk for metabolic disruptions later in life. For example, in the British Whitehall II Study, Kumari and colleagues (2004) found that low SES predicted higher incidence of diabetes relative to individuals with higher social standing. Using the same sample, Brunner and colleagues (1997) identified a link between SES and the metabolic syndrome. Similarly, in the Panel Study of Income Dynamics, a nationally representative prospective longitudinal study in the United States, Johnson and Schoeni (2011) found that poverty in childhood was predictive of diabetes almost 40 years later. In one of the few prospective studies of socioeconomic disadvantage and metabolic risk, Melchior, Moffitt, Milne, Poulton, and Caspi (2007) found that low SES in childhood was associated with a cluster of risk factors at age 32 (e.g., elevated blood pressure, adiposity, glycated hemoglobin). This finding remained significant, even when accounting for familial liability, childhood IQ, adolescent health behaviors, and adult SES. Recently researchers have incorporated sophisticated analytical techniques to separate the effects of early versus adult SES in the prediction of metabolic disorders. Using prospective data from the Health and Retirement Study, Nandi et al. (2012) used marginal structural models, a statistical approach that allowed them to estimate (a) the direct effect of early SES, (b) the direct effect of adult SES, (c) effects of early SES that are mediated by risk factors (e.g., blood pressure, smoking, alcohol consumption), and (d) the effect of early SES that is mediated by adult SES (referred to as a social trajectories model). By separately estimating these effects, Nandi and colleagues were able to test the relative contributions of early and adult SES on adult diabetes. Using this modeling approach, Nandi et al. found that early-life SES was associated with adult diabetes—an effect that was not accounted for by adult SES. Interestingly, when Nandi et al. used traditional regression analyses to compare the role of early and adult SES on later diabetes, the effects of early status were largely attenuated by

adult SES. In other words, comparison of the findings across the two statistical strategies suggests that studies that employ commonly used regression-based approaches for estimating effects (e.g., Agardh et al., 2007) may have significantly underestimated the effects of earlylife SES. Future studies should take advantage of these advancements in analytic techniques in order to more accurately estimate the relative contributions of childhood and adult risk factors. Although these studies offer convincing evidence for a link between low SES and metabolic abnormalities, recent evidence suggests that certain protective factors might play an important role in buffering children from the deleterious effects of early-life adversity. For example, low SES in childhood was associated with higher prevalence of metabolic syndrome at midlife in the MIDUS sample (Miller, Lachman, et al., 2011). However, for individuals who recalled high levels of maternal support, low SES was not associated with metabolic syndrome at midlife. Interestingly, this pattern was not explained by upward social mobility, suggesting that maternal nurturance, and not the addition of material resources, served as a buffer against low SES. These findings are encouraging, as they suggest that early adverse experiences might be offset by protective factors; continued examination of protective factors will be an important area for future research. Socioeconomic Disadvantage and Cancer A small but growing body of evidence suggests that early-life socioeconomic disadvantage may play a role in the development of cancer in adulthood (e.g., Lawlor, Sterne, Tynelius, Smith, & Rasmussen, 2006; Power, Hyppönen, & Smith, 2005; Pudrovska, Anishkin, & Shen, 2012). In the Wisconsin Longitudinal Study, which prospectively studied individuals over the course of 50 years, low early-life SES was associated with breast cancer. Similarly, in a prospective study of women in Great Britain (Power et al., 2005), trend-level associations between childhood social class and lung and stomach cancers emerged, even after controlling for health behaviors and adult social class. Additional research will be important to add further insight into the extent to which low SES in childhood is a predictor of cancer diagnosis in adulthood. Summary of Research on Early Socioeconomic Disadvantage and Later Disease The studies of socioeconomic disadvantage on later health outcomes have notable strengths compared with the maltreatment literature. They make use of large and representative samples across countries, use prospective designs, often have objective indicators of SES, and measure outcomes through medical records and health care databases (rather than self-report measures). Many of these studies control for plausible alternative differences in lifestyle factors, such as smoking, diet and exercise, and adiposity. The consistency of findings across countries suggests some universalism in the phenomenon and makes it unlikely that culture-specific factors (e.g., class-related differences in access to health care) are driving the findings. Despite these more rigorous study designs, the literature linking early SES with adult physical health has some weaknesses that complicate interpretation and conclusion. For instance, these studies largely make use of observational data, which is understandable, given that individuals

cannot be randomly assigned into poor environmental conditions. Further, the possibility remains that the observed links are due to unmeasured confounds, such as a common genetic liability that predisposes offspring to low SES and poor health. An additional limitation is that most of the studies are not fully longitudinal in that they do not assess childhood health during exposure to low SES (Cohen et al., 2010). Given that many of these disease endpoints (e.g., cardiovascular disease, stroke) are not observed in young children, assessment of the same health measures would not add incremental value to the study designs. Nonetheless, assessment of general physical health would help address the possibility that poor children start out life sicker than more affluent children, affecting both their parents' earning potential and their own long-term health. A handful of studies have capitalized on naturally occurring fluctuations in societal economic conditions to examine the impact of early-life resources on later physical health outcomes (e.g., van den Berg, Doblhammer, & Christensen, 2009; van den Berg, Doblhammer-Reiter, & Christensen, 2011). One advantage of these naturalistic experiments is that researchers can be fairly confident that a genetic liability is unlikely to be the cause of any observed connections between the adverse experience and physical health outcomes. Further, these studies still can be useful in documenting links between adverse experiences and later health problems, particularly when there are children from the same cohort who were not directly affected by the event (e.g., Heijmans et al., 2008). Moreover, given that the unusual event is due to random and external forces, these studies minimize the likelihood that unknown spurious factors (e.g., personality traits) explain the link between early exposure to adversity and later physical health problems. For instance, children who were born under adverse economic conditions in Denmark between 1873 and 1906 had higher mortality rates from cardiovascular disease than their peers who were born during more prosperous economic cycles (van den Berg et al., 2011). Further, van den Berg and colleagues (2011) reported that individuals who were born during a recession (and survived until at least age 40) lived, on average, 11 months less than other individuals born during an economically prosperous time, providing evidence that economic circumstances early in life may have a lasting effect on health outcomes. Not all studies of fluctuating economic conditions and long-term physical health have found support for this hypothesis, however. Cutler, Miller, and Norton (2007) did not find evidence of a link between economic fluctuations and physical health outcomes later in life in a sample of Americans who were born during the Great Depression. The detailed analysis of the Danish sample might shed light on why links between early-life economic circumstances and longterm health were not found in the U.S. study. In Denmark, the negative impact of economic adversity varied as a function of the community size: It was substantially less dramatic in major cities, such as Copenhagen, relative to smaller towns, where fewer resources were available. Similarly, individuals who lived in rural areas—with smaller communities of individuals who might take a more proactive approach to caring for each other—were similarly less affected than individuals who lived in midsize towns. It may be that a lack of early-life economic resources can be offset by community-level support systems, just as earlylife socioeconomic adversity within the family appears to be mitigated by maternal support

(e.g., Miller, Lachman, et al., 2011). Cutler et al. (2007) did not examine whether the link between early economic adversity and adult health varied based on the degree of urbanization in U.S. communities, which may have obscured these possible community effects.

Other Forms of Adversity and Later Disease To date, the epidemiological literature has provided some evidence that maltreatment and socioeconomic disadvantage may set the stage for a range of health problems in adulthood. Much less is known, however, about the extent to which other forms of adversity in childhood are risk factors for chronic diseases in adulthood. Of course, maltreatment and socioeconomic disadvantage are not the only forms of adversity that many children face that may shape the course of their long-term health trajectories. Next we briefly review evidence from several studies that have identified connections between other childhood adversities and later health outcomes. We note that these findings should be viewed as preliminary and will require further investigation in large, representative samples. Parental Death and Later Disease The death of a parent—resulting in the loss of an attachment figure—is widely regarded to be a profound stressor for children (e.g., Bowlby, 1980; Luecken & Roubinov, 2012). Studies have documented the risk for immediate and long-term mental health problems following parental death (e.g., Kivelä, Luukinen, Koski, Viramo, & Kimmo, 1998), but only recently have investigations studied the physical health consequences associated with parental loss in childhood. In an epidemiological study of women, Jacobs and Bovasso (2000) found that women who experienced the death of their mother in childhood were at increased risk for breast cancer. In another study, Krause (1998) examined early parental loss as well as recent life stressors in a national study of older adults. Although early parental loss alone was not associated with later disease, the interaction between parental loss and current stress levels predicted self-reported health and reports of acute conditions, including cancer, diabetes, and cardiovascular disease. In this case, the loss of a parent served as a vulnerability factor, and the addition of a current stressor increased the likelihood that individuals would have poor health outcomes. As we discuss in the next section, diathesis–stress models can add insight into links between early adversity and physical health, and they can shed light on specific contexts (e.g., stressful periods of life) and specific individuals (e.g., those who lost a parent in childhood) who might be especially prone to early onset of disease. Other studies, however, have not found links between early parental loss and later physical health (e.g., Maier & Lachman, 2000). When main effects do not emerge, it may be informative to examine whether there might be moderating factors that specify which individuals are most at risk, as Krause (1998) found. Interparental Violence and Later Disease To date, most studies examining the aftereffects of interparental violence on children's longterm outcomes have focused on emotional and behavioral problems (e.g., Hammen, Henry, & Daley, 2000), and little is known about how these early experiences might be associated with

long-term physical health problems. The large-scale Adverse Childhood Experiences study (Felitti et al., 1998) included interparental violence in the cumulative risk assessment for adverse childhood experiences, but they did not examine the unique role that interparental violence played in contributing to long-term physical health problems. Several studies have reported links between individuals' exposure to interparental violence as children and selfrated health problems in adulthood. In a large-scale European sample of adults (Roustit et al., 2011), the odds of having poor self-reported physical health were 2.3 times greater if participants reported exposure to interparental violence in childhood. This preliminary evidence suggests that exposure to interparental violence in childhood may be a risk factor for the early onset of disease in adulthood, but additional research is needed to better understand the role of interparental violence in contributing to later disease. Further, studies that incorporate disease diagnosis rather than self-reports of health will be helpful for establishing support for a link between intimate partner violence and chronic disease. Evidence from the study of children's short-term health outcomes associated with exposure to severe interparental conflict adds support to the idea that interparental violence may also be a risk factor for chronic health problems in adulthood (see Troxel & Matthews, 2004, for a review). Like much of the research on links between maltreatment and later physical health, this literature relies heavily on cross-sectional study designs and self-reports of physical health. Nevertheless, the evidence is compelling enough to warrant further examination of the role of interparental conflict in violence in contributing to long-term physical health problems later in life. Extended Parental Absence and Later Disease Some children encounter periods when their parents are absent for an extended period of time (e.g., military families where parents are deployed, families where a parent is imprisoned). Such long-term separations from parents in childhood have been associated with long-term mental health problems, such as depression (Pesonen et al., 2007). Less is known about how these extended separations might be linked to later disease outcomes. Some naturalistic evidence from World War II suggests that long-term parental separation may be associated with physiological stress activity in late adulthood (Pesonen et al., 2010). During World War II, approximately 70,000 children were separated from their parents and sent to live in foster care to escape the dangers of war. Many of these children were separated from their parents for several years. Pesonen and colleagues (2010) found that adults who had been separated from both parents as children had higher average salivary cortisol and greater salivary cortisol reactivity during a laboratory stress task relative to adults who had not experienced long-term parental separation in childhood. Given the large numbers of children who experience lengthy parental separations due to ongoing military deployments and divorce, there is a need to examine what impact, if any, such disruptions in the parent-child relationship will have on long-term health outcomes. Natural Disasters and Later Disease Natural disasters, including hurricanes, tsunamis, tornadoes, earthquakes, and drought, result in

dramatic disruptions to daily life. These disruptions range in severity from temporary setbacks to major transformations, including displacement and loss of one's home, inadequate access to food and other basic necessities, and uncertainty about the future. In some ways, natural disasters differ from exposure to some of the other adversities that have been a focus of this chapter, in that the cause of the distress is often not traceable back to poor parental care or disparities in resources. At the same time, however, the aftereffects that could result from living through a natural disaster may resemble the consequences of stressors that co-occur with poverty and maltreatment. Several studies have documented the ways in which physical health can be affected in the short term by exposure to natural disasters. For example, in 1995, Japan experienced a 6.8magnitude earthquake near Kobe, resulting in the loss of over 6,000 lives and leaving over 300,000 homeless. Following this disaster, there was a threefold increase in heart attacks and twofold increase in strokes, particularly for individuals who lived close to the epicenter (Kario, McEwen, & Pickering, 2003). These findings offer compelling evidence suggesting that natural disasters create enormous stress for individuals that might have serious health implications. To date, however, evidence for the long-term health impacts for individuals who endured natural disasters as children has been lacking. Further, although evidence suggests that children experience physical health problems in the months following a natural disaster (Datar, Liu, Linnemayr, & Stecher, 2013), the long-term effects of these stressful experiences still are largely unknown. War and Later Disease The immediate effects of war cause significant stress on children and their families. Children may face daily threats of violence, loss of loved ones, and fears about the future. In some cases, the effects of war have long-ranging influences on daily life, including reduced access to resources and basic necessities. For example, during World War II, a German blockade cut off the food supply to a large area of the Netherlands during the winter of 1944–1945. This tragic event provided the rare opportunity to examine the connection between starvation during pregnancy and long-term health outcomes of offspring affected by the war-induced famine (de Rooij, Wouters, Yonker, Painter, & Roseboom, 2010). Exposure to famine and malnutrition prenatally has been linked to Type 2 diabetes, cardiovascular disease risk, other metabolic abnormalities, and even breast cancer in adulthood (Barker & Clark, 1997; de Rooij et al., 2010; Painter et al., 2006; Painter, Roseboom, & Bleker, 2005; Rich-Edwards et al., 1999). Findings from survivors of World War II suggest that exposure to wartime stressors may be associated with elevated rates of cancer. In one study, cancer rates were elevated in Europeanborn immigrants who immigrated to Israel before or during the war (Keinan-Boker, Vin-Raviv, Liphshitz, Linn, & Barchana, 2009). The largest effects emerged for people born between 1940 and 1945, who would have been exposed to horrific conditions before age 5. Their cancer risk was elevated 3.5-fold relative to same-aged immigrants who arrived in Israel before the war. Like the experience of socioeconomic disadvantage and maltreatment, these adversities described are likely to expose children to chronic, stressful circumstances, with lingering

threats of repeated exposure to the stressor (e.g., lasting concerns of repeated interparental violence) and aftereffects resulting from the stressor (e.g., loss of all resources as a result of a natural disaster). To date, these adversities have not been examined extensively in relation to long-term health outcomes. As we develop a better understanding of the ways in which early adversity predicts later health outcomes, it will be important to understand the characteristics and contexts of early adversity that are most likely to contribute to poor health. Increased specificity about the types of adversities that influence health will help researchers and policymakers focus efforts on mitigating the risks that are most likely to contribute to chronic disease.

Limitations and Alternative Explanations Although the evidence linking early adversity to physical health is provocative, there are a number of weaknesses to this research that preclude a definitive interpretation. For example, many of the studies rely on one-time assessments, during which early stress and later health are measured via retrospective self-report. Understandably, this methodology raises concerns about the reliability of reporting, which would be particularly problematic if reporters are systematically biased in their assessments of early adversity and adult health. Some evidence suggests, however, that adults underreport negative experiences from childhood (Dill, Chu, Grob, & Eisen, 1991). Another concern is that these studies may reflect differences in the reporting of health problems as a function of early adversity, rather than real differences in the manifestation of disease. Notably, however, many studies—particularly studies linking socioeconomic disadvantage to later health outcomes—utilize medical records and physician diagnoses, thereby eliminating informant bias in symptom reports (e.g., Felitti et al., 1998; Rich-Edwards et al., 2012). Yet another possibility is that genetic confounds might play a large role in creating the observed connections between early adversity and physical health. For example, there could be a familial genetic liability that contributes to abusive parenting and disease vulnerability. Or it could be the case that the direction of effects is wrong. For instance, a sick child can accumulate enormous medical bills, not to mention chronic stress on caregivers and other family members. The net effects of expensive medical care and lost earnings due to time spent caring for sick children might result in low SES in adulthood. A handful of studies have controlled for child sickness (e.g., Caspi et al., 2002), but this practice needs to be adopted more widely before we can definitively rule out this alternative explanation. Using Animal Studies to Supplement Epidemiological Evidence In light of the fact that it would be unethical and impractical to experimentally manipulate early human experiences or exposure to pathogens in adulthood, it is difficult and perhaps impossible to evaluate whether early adversity causally influences adult disease in humans. However, studies of this nature can be done with animals, and it is useful to consider how these studies have informed our understanding of connections between early adversity and adult physical health. In this section, we describe some of the findings that have emerged from experimental studies with animals that bolster support for the findings from epidemiological

studies with humans. Evidence from studies using experimental manipulations with animals suggests a causal link between early adversity and health outcomes. Animal models often provide good mechanistic evidence relating psychosocial risk factors to physical health. Unlike humans, animals can be randomly assigned to psychologically difficult circumstances and monitored over time to determine whether these stressful experiences influence disease outcomes. Further, questions about the effects of the relative timing of adverse experiences (e.g., early infancy versus later childhood) can be addressed by manipulating the onset of the stressor and examining differences in later health consequences associated with these timing differences. Often findings from animal studies converge with correlational evidence from human epidemiologic research. Hofer's (2002) observations that rat pups responded with significant distress when separated from their mothers—similar to distress signals in humans—raised questions about the mechanisms of separation distress in nonhuman mammals. To explore this question, Hofer conducted a series of experiments to identify what physiological subsystems become disrupted when mothers were removed from their pups (for a review, see Hofer, 2002). These studies revealed a number of hidden regulators that no longer functioned properly if rat pups were separated from their mothers. Hofer found that when rat pups were removed from their mothers, they showed a dramatic decline in multiple physiological and behavioral systems, such as those controlling heart rate, body temperature, food intake, movement, and exploratory behavior. Hofer argued that maternal separation resulted in the removal of the important regulatory functions that the mother serves for her offspring, and the absence of these regulatory components resulted in rat pups' visible behavioral distress. These studies led Hofer to conclude that mother-infant interactions have embedded within them a number of vital physiological regulatory functions that are negatively affected by separation from maternal care. In this regard, animal studies have some parallels to the experiences of children in poverty or unsupportive caregiving environments. On the whole, children exposed to these stressors receive less sensory, cognitive, and emotional stimulation and are more likely to be deprived of basic necessities, such as food and heat, than children growing up without extreme adversity (e.g., Evans, 2004). That said, it is difficult to know how closely these deprivation studies in animals actually resemble human experience. Permanent maternal separation and extreme sensory deprivation are relatively uncommon in humans, even in families where maltreatment and disadvantage are present. At best, what these animal studies model is extreme and unusual human stress. Species differences in developmental growth are another major challenge in translating conclusions from animal studies to meaningful information for human development. At the time of birth, rodents are much less mature than humans. As a result, their physiology may need external regulation by caregivers in a manner that is not comparable to full-term human newborns. These caveats aside, a number of studies indicate that premature separation can have long-term effects on animals' susceptibility to disease. In one study, mice pups were separated from their

mothers for 6 hours a day over the first 2 weeks of life (Avitsur, Hunzeker, & Sheridan, 2006). As adults, the mice were challenged with intranasal exposure to an influenza virus. Compared to controls that remained with their mothers until weaning, the separated mice had greater viral replication and worse symptoms of infection, which was a result of an overly aggressive inflammatory response to the virus. Moreover, the proinflammatory response was present 9 days after the infection, a time when viral particles had declined to the point of being almost undetectable. These findings suggest that early stress calibrated the immune system to mount overly aggressive and extended inflammatory responses to the virus. In another study, rats endured a stressful situation at 100 days of life (corresponding to early adulthood in humans). In order to obtain food and drink over a 4-day period, rats incurred electric shocks (Ader, Tatum, & Beels, 1960). Nearly all of the rats developed gastric ulcers following this stressful experience. However, the density of these ulcers was fivefold greater in rats that had been prematurely separated from their mothers at 15 days of life. In a follow-up study, a third group was added to evaluate whether nutritional disparities accounted for the health effects of the maternal separation manipulation. To test this hypothesis, rat mothers were removed from her offspring at 15 days, had their nipples cauterized to prevent nursing, and then were returned until 21 days. The median number of ulcers in this group was similar to the control condition. Because both groups in which the mother remained present had significantly fewer ulcers than the prematurely separated and weaned rats, these findings suggest that the effects were due to the absence of maternal nurturance rather than nutritional deficiencies per se (Ader, 1962). There also has been mounting interest in early-life influences on asthma in animal models. In one study, mice were randomized into one of three conditions: In one condition, they received regular footshocks for 1hr on 3 days during the fourth week of life; in another condition, mice watched and heard other mice undergo this experience but were not shocked themselves; and in the third condition, the mice were undisturbed in their home cages (Chida, Sudo, Sonoda, Hiramoto, & Kubo, 2007). When the mice reached young adulthood (at 8–10 weeks of life), they were sensitized to ovalbumin, a protein in eggs that causes allergic reactions. At 11 weeks, all mice were given airway challenges with ovalbumin. Relative to controls, those mice that received or observed footshocks showed greater airway inflammation and more bronchial reactivity to the challenge. Similar patterns were observed in another study of rats that, over the first month of life, were either separated from their mothers daily for 2 hours and then reunited or remained undisturbed in their home cages (Kruschinski et al., 2008). When the rats were adults, asthma was induced by sensitizing subjects to ovalbumin, and airway tissue was collected. Analyses revealed striking differences. Adult rats that had been repeatedly separated from their mothers early in life showed more severe airway pathology than adult controls, with increased numbers of eosinophils, T-cells, and other proinflammatory mediators found in their lungs upon dissection. Summary Collectively, these studies provide evidence to support the notion that early adversity has longterm effects on physical health. To be sure, more work needs to be done to definitively rule out

alternative explanations and clarify what the associations reflect in terms of underlying pathophysiology. But the weight of the evidence, from observational data and natural experiments in humans and randomized studies in animal models, leads us to conclude that there is evidence indicative of a causal effect. With that said, increased precision in the definition and assessment of early adversity and adult health will aid in our understanding of how and under what conditions early adversity leads to later health problems. This increased specificity in measurement will provide insight into what characteristics and contexts of early adversity prove to be most detrimental for physical health outcomes. Further, increased attention to whether there are protective factors that can mitigate the risk associated with early adversity and what mechanisms put people on a trajectory from early adversity to poor health across development will be an important direction for future research.

Conceptual Models Linking Childhood Adversity to Adult Physical Health Researchers have long recognized that no single risk factor can account for all chronic diseases of aging, and research on the early origins of adult physical health outcomes has supported the notion that there are multiple pathways linking early experience to adult disease (e.g., Felitti et al., 1998; Miller et al., 2011; Taylor, Repetti, & Seeman, 1997). Yet to date, much research on the experience of early adversity and adult health has explored risk factors in isolation without consideration of the broader context in which the risk factors occur. Moreover, despite an interest in the impact of early adverse experiences on later health, existing studies have not thoroughly examined the relative timing of adverse experiences. Evidence from other areas of developmental research suggests that meaningful critical and sensitive periods constrain the impact of early experience on later outcomes (e.g., Knudsen, 2004), and it stands to reason that similar sensitive periods may constrain the impact of early adversity on long-term physical health outcomes as well. Given that chronic disease is a multiply determined phenomenon, it is useful to clarify the complex pathways from childhood adversity to adult physical health by testing larger conceptual models that integrate multiple risks across multiple time points. In this section, we review several theoretical models that can be used to guide the investigations of connections between early-life adversity and adult physical health. Many of these models take into account the notion that multiple factors contribute to physical health outcomes; further, these models describe processes through which intervening factors (e.g., buffers, vulnerability factors) can alter individuals' health trajectories over time, leading to endpoints that may not have been expected, given individuals' starting points. These models provide a foundation for clear, testable hypotheses about the processes through which early adversity contributes to disease risk.

Cumulative Models In contrast to models that examine a single risk factor in isolation, cumulative risk models

provide a testable framework for studying how risk factors operate in the context of other risk factors to influence later health outcomes. Cumulative models of risk focus on the accumulation of risk factors and on how an increase in the number of risk factors an individual is exposed to might translate into an increase in risk for poor functioning (Sameroff, 2000). In other words, cumulative models predict that as the number of risk factors increases, there is an increase in the likelihood of poor outcomes. Two types of cumulative risk models, described next, offer variations in the hypotheses about how the presence of additional risk factors might be associated with poor physical health outcomes in adulthood. Linear Risk Models One type of cumulative risk model is a linear model, which emphasizes that each increase in early risk is associated with significantly greater risk for later disease. Two key assumptions go along with this model. First, one must assume that each adverse experience conveys more or less equal risk in shaping health. Linear models do not place any weight on the relative severity of risk factors. Instead of focusing on the characteristics of the risks, linear models focus on the quantity of risk factors. The second assumption of this model is that the various adverse experiences are nonredundant in their ability to predict later health. In other words, each risk factor creates additional strain or burden on individuals in a way that leads to worse health in adulthood. Although some early adverse experiences, such as neglect and poverty, may lead to the same specific deficit in access to resources (e.g., poor or neglected children may struggle to receive adequate food and health care), according to a linear risk model, an individual who is poor and neglected would have worse health outcomes than an individual who is poor but not neglected. Support for a linear risk model comes from a recent study by Evans and Kim (2012), who examined the role of cumulative risk exposure as a predictor of markers of physiological wear and tear (i.e., allostatic load, a biological marker described in more detail in the next section). In this study, researchers created an index of cumulative risk exposure score by summing across items about physical risk (e.g., noise, housing problems) and psychosocial risk (e.g., violence in the home). Evans and Kim found that cumulative risk exposure at age 13 was a predictor of allostatic load at age 17. These findings are particularly notable because the markers of wear and tear (including measures of endocrine, cardiovascular, and metabolic functioning) were measured while participants were still young (i.e., during adolescence), so it will be important to examine whether cumulative risk exposure predicts individuals' accelerated aging and chronic disease later in life. Threshold Risk Models Threshold risk models, like linear risk models, focus on the number of risk factors present. Unlike linear risk models, however, threshold models predict that only after people experience a certain number of adverse experiences will their risk for poor health increase. Once people pass this threshold, they are likely to experience health problems. In this way, threshold models allow for the experience of some adversity without necessitating that each exposure to stress must lead to later health problems, as is the case in linear models. In other words, for every

health outcome, there is some tipping point at which early adversity poses a serious risk for poor health. Moreover, at this tipping point, risk factors may have synergistic properties, such that the combined effect of the risk factors is much worse than the sum of each risk factor in isolation. Some of the best evidence for threshold risk models comes from the Adverse Childhood Experiences study, which examined over 9,000 individuals who were enrolled in a large insurance health maintenance organization (Felitti et al., 1998). Participants completed selfreport assessments about 17 adverse experiences that may have occurred in childhood, including abuse, violence, and parental mental illness. These experiences were categorized into seven domains of adversity, and participants received a total score indicating the total number of domains in which they were exposed to risk. This cumulative risk score was then used to predict physician-diagnosed diseases. As predicted, early adversity was associated with heart disease, cancer, and strokes. Interestingly, however, individuals were at increased risk for these diseases only when they reported four or more adverse experiences. When participants reported three or fewer adverse experiences, they were not at increased risk for these diseases. Notably, in many studies, adverse experiences cluster together more than one would expect by chance (Crouch et al., 2000). This pattern of clustered risks may explain some of the observed threshold findings. For example, one risk in isolation might reflect circumstances that are beyond the control of caregivers (e.g., single parents in poverty may have limited abilities to raise their social status). Children who experience poverty but not maltreatment, parental psychopathology, or other serious risks may grow up in an environment that is lacking in material resources but otherwise stable and conducive to healthy development. These children may in fact be shielded from some of the most toxic aspects of poverty. Despite their economically stressed conditions, they would have access to a supportive caregiver who is reliable and responsive to their needs. In contrast, children who are exposed to many adverse experiences might endure chaotic, unsupportive family contexts that are ill-equipped to meet the needs of growing children. For these children, who face a multitude of risks, it is quite possible that their total exposure to stress is well beyond the sum of each stress alone; such multiplicative effects can be identified with threshold models. In summary, both linear and threshold cumulative models offer a heuristic for explaining the role of early adversity in shaping later health. These models suggest that risk is best captured at the level of total exposure to risk, with the general framework that greater exposure to adversity will translate to more physical health problems in adulthood. These models do not differentiate between different domains of risk exposure (e.g., abuse from primary caregivers, neighborhood violence, access to resources); instead, these models posit that adverse experiences across domains will exert a negative influence on long-term health outcomes. In addition, these models typically do not take into account the timing of exposure to adversity, which limits researchers' ability to examine whether there are sensitive periods in which the experience of early adversity is most likely to translate into health problems in adulthood. Further, cumulative models explicitly ignore the relative severity of risk factors; as such, these models can be viewed as relatively crude approaches to defining risk. Yet despite this

somewhat unsophisticated approach to quantifying adversity, many researchers have found strong evidence for cumulative risk models in the prediction of chronic disease outcomes in adulthood.

Diathesis–Stress Models One question of great interest to developmental psychopathologists focuses on why some individuals seem to be at greater risk than others for poor functioning (Shonkoff & Phillips, 2000). One popular approach to answering this question is through testing of diathesis–stress models (e.g., Ingram & Luxton, 2005). In these models, a particular risk factor (e.g., poverty) is associated with poor outcomes only for individuals who have some sort of vulnerability, which might include a genetic risk factor, temperamental quality, or other personality characteristic. These vulnerability factors increase individuals' susceptibility to the negative effects of adversity. This model can be useful in examining connections between early exposure to adversity and adult health by focusing on whether there are specific individuals for whom stressors might be most likely to lead to poor health outcomes as a result of their early experiences. To date, most diathesis–stress studies of vulnerability to early adversity have focused on adult mental health outcomes (e.g., Moffitt, Caspi, & Rutter, 2006). Recent evidence suggests that diathesis–stress models will be useful in future examinations of the links between early adversity and physical health. Brody and colleagues (2013) examined the family emotional climate in early adolescence and two gene polymorphisms (the short allele of the 5-HTTLPR gene and 7+ repeat of DRD4) as predictors of allostatic load in early adulthood. These two genetic polymorphisms were of particular interest because they have been suggested to confer sensitivity to family environments (e.g., Bakermans-Kranenberg & van IJzendoorn, 2011; Belsky, Bakermans-Kranenberg, & van IJzendoorn, 2007). Brody et al. found that individuals who had the short allele for 5-HTTLPR and 7R allele for DRD4 and who were exposed to less supportive family environments had greater allostatic load (a composite measure that included body mass index (BMI), blood pressure, and neuroendocrine functioning) relative to individuals without both “sensitivity” alleles. Individuals with these same alleles but living in supportive family environments did not have elevated allostatic load. Further, the quality of the family environment was not directly associated with allostatic load. These findings are provocative in suggesting more nuanced ways that family experiences might confer risk for allostatic load for certain individuals with particular genetic vulnerabilities. This line of research is still in its infancy; additional work is needed to better understand what vulnerability factors (e.g., genetic liabilities, temperament) best identify individuals for whom early adversity is most likely to be associated with poor health in adulthood.

Differential Susceptibility In contrast to diathesis–stress models, models of differential susceptibility posit that vulnerability sources actually are better viewed as plasticity factors that magnify the risk for poor outcomes in adverse conditions and amplify the likelihood of positive outcomes in more supportive environments (Belsky, 1997; Belsky et al., 2007; Ellis, Boyce, Belsky, Bakermans-

Kranenberg, & van IJzendoorn, 2011). Theories of differential susceptibility, including Belsky's (2005) differential susceptibility theory and Boyce and Ellis's (2005) biological sensitivity to context theory, are similar to diathesis–stress models in that they predict that certain individuals with particular characteristics (e.g., genetic predispositions, temperaments) will be more negatively affected by unsupportive environments relative to other individuals who have different genetic or temperamental characteristics. At the same time, however, differential susceptibility models assert that the traits that put some individuals at risk for exponentially worse outcomes also predispose the same individuals to better-than-expected outcomes given supportive contexts, leading researchers to argue that these individuals are both “for better and for worse,” depending on their environment (Belsky et al., 2007, p. 300). Like diathesis–stress models, differential susceptibility models may be useful as a way of documenting why some individuals are more adversely affected than others by exposure to harsh conditions in early life. Differential susceptibility models take this idea a step further, however, by suggesting that the same individuals who flounder in adverse conditions may be ones who would thrive in more supportive contexts. Evidence for differential susceptibility has been found for the prediction of a variety of outcomes, including mental health and behavior (for a review, see Belsky & Pluess, 2009), but to our knowledge, no evidence has been documented for differential susceptibility in studies of early adversity and adult physical health. Notably, however, it was Boyce et al.'s (1995) observations of children's illness rates that motivated the theory of biological sensitivity to context. In this study, highly reactive children had the lowest rates of illness in supportive environments and the highest rates of illness in unsupportive environments. Thus, it seems reasonable to predict that similar models of differential susceptibility will be evident in the study of early adversity and adult physical health.

Buffering Models Similar to diathesis–stress models, buffering models help clarify the specific contexts in which risk factors are associated with poor outcomes. The main difference, however, is that buffering models uncover protective factors that can help mitigate the negative effects associated with early risk (e.g., Chen Miller, Kobor, & Cole, 2011; Miller, Lachman, et al., 2011). These models offer great starting points for the development of interventions; once a protective factor has been identified, interventions can target this domain with the goal of minimizing the negative outcomes associated with risk. For example, as mentioned earlier in this chapter, Miller and colleagues (2011) examined the connection between early SES and metabolic syndrome at midlife and found that maternal care served as a buffer, such that individuals from poor backgrounds who experienced high levels of maternal nurturance were not at risk for metabolic syndrome. Given that many of the early adversities that children face are difficult to resolve quickly (e.g., poverty is a chronic and pervasive form of adversity), efforts to uncover buffers that can lessen the negative impact of adverse experiences may prove to be particularly instrumental from a public health perspective.

Developmental Trajectory Models One of the core themes of development is the notion that individual differences exist across development (Shonkoff & Phillips, 2000). Waddington (1957) argued that early in development, organisms have a range of possible developmental pathways (what Waddington called the epigenetic landscape). Over time, these pathways become more constrained and deterministic as a function of increasing constraints and limitations on the developing system. Further, the developing behaviors and systems become more canalized and resistant to change over time. According to Waddington, changes in the developmental trajectory are most likely to occur at decision points—transition periods with multiple options for the subsequent course of development. This theory proposes that major deviations from normative development early in life (e.g., child neglect) would lead to major changes in expected outcome (in this case, physical health) relative to minor deviations that take place at later points in development. Waddington's ideas about developmental pathways became the foundation for many theories that sought to explain variation in developmental trajectories. Developmental trajectory models posit that there is a progressive pruning of possible pathways across development, an idea that is consistent with other developmental theories that emphasize the importance of early critical or sensitive periods in development (e.g., Bowlby, 1973; Shonkoff & Phillips, 2000). The issue of sensitive periods has been relatively unexplored in relation to early adversity and adult physical health. There is an assumption that early childhood represents a sensitive period in which adverse experiences have a disproportionately large influence on health trajectories, but empirical tests of this idea have been lacking. Is early childhood a unique time for the development of long-term health problems? Might there be other periods of development—such as adolescence—in which exposure to adversity is likely to confer greater risk for health problems? Some evidence suggests that each exposure to adversity at different developmental periods confers risk for poor health outcomes (e.g., Singh-Manoux et al., 2004). Other evidence, however, suggests that experiences in infancy and early childhood are particularly potent for the long-term risk for infectious disease (e.g., Cohen, Doyle, Turner, Alper, & Skoner, 2004). For example, Cohen et al. (2004) examined adult participants' reactions to virus exposure as a function of early-life socioeconomic conditions. Following intranasal exposure to rhinovirus, participants were monitored for cold symptoms. As hypothesized, childhood SES was associated with cold systems. Notably, the strongest effects were found for SES at the earliest ages of measurement (12–24 months), with a progressive decline in the strength of the connection between SES and susceptibility to colds when examining later-childhood and adolescent family SES. Evidence from some animal studies suggests that there may be windows of development in which stressful experiences lead to biological tendencies (e.g., Meaney & Aitken, 1985) and health outcomes (e.g., Ackerman, Hofer, & Weiner, 1975). These kinds of developmental timing studies are difficult to conduct in humans; intervention studies and the use of naturalistic experiments can help identify whether there are particular periods in development during which individuals are especially sensitive to exposure to adverse conditions.

Waddington's epigenetic landscape can be used to explain the processes involved in two developmental trajectory models (described in the next two subsections). These models document the processes through which development unfolds over time in a nonlinear pattern and can help explain how (a) individuals with different backgrounds and risk exposure experience the same physical health problems later in life and (b) how individuals with the same risk exposure ultimately experience dramatically different outcomes in adulthood. These models can serve as useful guides for understanding the complex ways in which early childhood adversity can contribute to adult physical health. Equifinality As we have already discussed, there is evidence linking early adversity to cardiovascular disease, diabetes, and a number of other chronic diseases of aging. One question that arises when looking at the links between adversity and health is: How do different adverse experiences—many of which are nonoverlapping experiences—lead to the same poor health outcomes? What happens across development that leads individuals with different backgrounds ultimately to share the same chronic diseases? Models of equifinality document the process through which different starting points in development may lead to the same outcome (Cicchetti & Rogosch, 1996). As Waddington (1957) theorized, individuals will vary in their developmental trajectories, in part because of variations in choices and conditions at key decision points that shape subsequent experiences. As such, it is possible that a complex series of decisions made across the early years of life will set into motion a series of developmental trajectories that share a common endpoint (e.g., diabetes, cardiovascular disease). It may be that even though individuals from impoverished or abusive backgrounds experience different day-to-day stressors in childhood, the consequences of such stressful experiences may have the same effect on physiological regulatory systems, coping and regulation of emotional distress, and health behaviors that shape the course of disease. For example, early adversity—whether such adversity includes maltreatment or poverty—may cause individuals to be vigilant toward threatening information in the environment and mistrusting of others. These characteristics shape the manner in which people engage their social worlds, making them more likely to elicit conflict and rejection and less likely to garner warmth and support. The experience of early adversity may shape success in forming and keeping long-term, high-quality relationships. Early adversity also is associated with poor self-regulation and coping skills, wherein the future is highly discounted in favor of immediate gratification, including engagement in risky or unhealthy lifestyle behaviors (e.g., high-fat diets, excessive alcohol consumption, drug use). Over time, these proclivities and experiences exert wear and tear on the body, with potential repercussions for disease outcomes. Thus, even though two individuals may have had unique starting points and sources of adversity (e.g., poverty versus abuse), they may come to share the same chronic health problems in adulthood because of a series of events across the life span that constrained their development and shaped their progression toward disease (Cicchetti & Rogosch, 1996; Miller et al., 2011). Multifinality

Like models of equifinality, the focus of multifinality models is on the trajectories that explain developmental processes and how unique experiences can change the long-term path toward health or disease. Of course, not all individuals are affected in the same way by the experience of early adversity. Multifinality models help explain how individuals with similar starting points (e.g., growing up in poverty) develop varied outcomes over time (Cicchetti & Rogosch, 1996). These models can be particularly informative for understanding the different mechanisms that explain how individuals who share common adverse experiences in childhood face unique health problems decades later. Individuals can take different trajectories for a number of reasons. For example, intervening factors can offset negative trajectories, as illustrated in tests of buffering models (e.g., Miller, Lachman, et al., 2011). In these cases, individuals who experience adversity do not go on to develop poor health outcomes because something in their development (e.g., a supportive parent, a new role model) changed the course of their trajectory. Similarly, as discussed in models of diathesis–stress and differential susceptibility, individuals might possess vulnerability or plasticity factors (e.g., genetic liabilities, temperamental traits) that explain why some individuals are more adversely affected by negative environments than others (e.g., Brody et al., 2013). Further, evidence for multifinality can emerge when individuals who are exposed to the same adversity go on to face different disease outcomes in adulthood (e.g., heart disease versus cancers versus autoimmune disease). In this case, the nature of the disease endpoint may differ across individuals for a variety of reasons (e.g., genetic predispositions), but disease morbidity can be traced back to the same initial stressor.

Developmental Cascades Another model of development emphasizes the possibility that early experiences may influence the development of subsequent systems or domains (Masten & Cicchetti, 2010; Rutter & Sroufe, 2000). These developmental cascade models are particularly useful in depicting the connections between early experience and later adaptation (or maladaptation), particularly when the causal explanation for such a link is unclear. These models explain how problems in one domain of functioning (e.g., physical abuse in the parent-child relationship) can spill over into other areas of functioning (e.g., poor lifestyle behaviors) across development. Methodologically, researchers measure multiple domains of functioning across several time points (usually at least three of each; Masten, Burt, & Coatsworth, 2006). Analytically, cascade models control for the longitudinal stability of each domain while also including the cross-lagged paths across domains to test the extent to which one domain of functioning is influencing another. Developmental cascade models might provide insight into how it is possible that detrimental effects resulting from early experiences of childhood adversity might lie relatively dormant for decades, only to emerge as experiences of chronic disease in midand late adulthood. For example, the experience of adversity in childhood is unlikely to transform immediately into cardiovascular disease or other chronic diseases of aging. Nevertheless, as described earlier in relation to individuals' developmental trajectories, exposure to early stressors might influence children's health behaviors, interactions with others, and outlook for the future. These subtle intervening steps along the path from early

adversity to adult health would represent instances in which early exposure creates a cascade of aftereffects, which reverberate biologically across numerous systems and ultimately lead to chronic diseases in adulthood. In this way, cascade models can identify the steps along the progression from early adversity to later physical health problems. Although these models have become quite popular in identifying spillover effects associated with mental health problems and social functioning (e.g., Bornstein, Hahn, & Haynes, 2010), cascade models have been underutilized in investigations of early adversity and adult physical health, in part because longitudinal studies are needed with multiple waves of assessments of adversity and health. Another difficulty with these study designs is that the same measure of health across ages might be impractical. For example, disease outcomes such as diabetes and cardiovascular disease are less likely to emerge before adulthood, so equivalent health measures across time points may not be available. One study examined the cascading effects from family adversity and poverty in childhood to self-reports of health in adulthood (Herrenkohl et al., 2010). In this study, childhood family adversity was associated with adolescent internalizing symptoms, which in turn were associated with poor self-reported health in adulthood. These findings demonstrate that cascade models can be useful in identifying mechanisms that explain the progression from early adversity to later health. Research is needed using these types of models in samples with more objective measures of adult health.

Biological Intermediaries Linking Early Adversity to Adult Physical Health Now that links between early adversity and adult physical health have been clearly identified, the next step is to identify what biological processes translate the stressful experiences into the physical environment of disease pathogenesis (i.e., what biological mediators carry the effects of early adversity under the skin). In the last several decades, significant progress has been made in understanding the biological correlates of stress. This burgeoning area of research is due, in part, to Campbell and Fiske's (1959) long-standing calls to conduct multitraitmultimethod research, as well as directives to conduct research at multiple levels of analysis (e.g., cellular, molecular, individual, family, and societal levels; Cicchetti & Dawson, 2002). These data provide new insights and establish a conceptual approach for future studies mapping relations between early adversity and physical health. Yet at the same time, there are limits to what these biological mechanisms can tell us about why early adversity is an important determinant of physical health and disease. In this section, we provide a nonexhaustive review of several of the most widely studied biological processes thought to translate early adverse experiences into chronic diseases of aging. These biological mediators have received considerable theoretical and empirical attention for their potential role in explaining how early adverse experiences contribute to disease and accelerated aging. We provide a critical review, highlighting the strengths and weaknesses of each mechanism and documenting the limits of conclusions that can be drawn

when evidence is found for a link between biological processes and early adversity. We note that other biological mediators have been examined in relation to early adversity and later health outcomes (e.g., sympathetic and parasympathetic nervous system activity), but our focus is on mediators that have been, and continue to be, the major thrust of research in this area.

Hypothalamic-Pituitary-Adrenocortical Axis One potential mechanism that has received widespread attention is the HPA axis. This hormonal response system is present in organisms ranging from birds to humans and can be activated by a broad range of mental and physical stressors (McEwen, 1998; McEwen & Stellar, 1993; Weiner, 1992). Next we provide a review of the HPA axis and common measures used to capture HPA axis activity. After that, we review evidence supporting the notion that HPA axis activity functions as a biological mechanism linking early adversity to adult health outcomes. We then discuss some important caveats that should be taken into consideration when evaluating the role of HPA axis functioning as a mechanism linking early adversity with health outcomes. Cortisol: Background and Measurement Activation of the HPA axis occurs when neurons in the paraventricular nucleus of the hypothalamus secrete corticotropin-releasing hormone (CRH). This molecule travels to the anterior pituitary gland, which responds by secreting a pulse of adrenocorticotropin hormone (ACTH). The ACTH signal is carried in the blood to the adrenal glands, which synthesize and release cortisol into the bloodstream. Cortisol then travels throughout the body through the bloodstream and acts on a wide range of tissues. Of all the hormones released in this cascade, cortisol has gained the most notoriety, probably because of its widespread regulatory influences. Cortisol is a steroid hormone belonging to the glucocorticoid family, and it is thought to play a key role in the biological stress response system. In addition, cortisol plays an important role in the central nervous system, where it is involved in learning, memory, and emotion; in the metabolic system, where it regulates glucose storage and utilization; and in the immune system, where it regulates the magnitude and duration of inflammatory responses and the maturation of lymphocytes (Sapolsky, Romero, & Munck, 2000). Cortisol's role in regulating such a wide range of physiological systems has helped motivate research focused on its possible role in linking early adversity to adult physical health. Cortisol levels in the body exhibit a diurnal rhythm, characterized by high levels upon waking, proceeded by a rapid increase in cortisol after approximately 30 minutes (a feature known as the cortisol awakening response [CAR]), followed by a progressive decline in cortisol levels across the day (Kirschbaum & Hellhammer, 1989). Thus, cortisol research tends to incorporate one or more of the following indices of cortisol: (a) the size of the CAR; (b) the slope of the cortisol diurnal rhythm; (c) reactivity to naturalistic or laboratory-based stressors; and (d) metrics reflecting total cortisol output over some unit of time, typically the day, as reflected by area under the diurnal curve or urinary cortisol output (Adam & Kumari, 2009; Miller, Chen, & Zhou, 2007). These measures of cortisol output are thought to provide distinct information about HPA axis functioning. For instance, the CAR, which is regulated in part by the

suprachiasmatic nucleus (Clow, Thorn, Evans, & Hucklebridge, 2004), is thought to play a role in mobilizing energy stores for the demands of the day (Hucklebridge, Mellins, Evans, & Clow, 2002). Further, an increase in the CAR is thought to reflect an increase in life stress, whereas a reduced or blunted awakening response is indicative of a failure to recruit the resources needed to tackle the day's demands (Chida & Steptoe, 2009). Normatively, cortisol has a diurnal rhythm that is characterized by a pronounced decline from morning through evening, although there appears to be substantial intra- and interindividual variability in this pattern, even among healthy persons (e.g., Stone et al., 2001). Nonetheless, flat rhythms that lack the typical diurnal decline often are seen in persons exposed to chronic stress (Gunnar & Vazquez, 2001) or with chronic psychiatric disorders, such as posttraumatic stress disorder (e.g., Yehuda, Golier, & Kaufman, 2005). Indices of total cortisol output are thought to provide a cumulative estimate of how much of this hormone bodily tissues get exposed to over some defined period of time, such as a day. Levels that are either too high or too low are thought to be detrimental for health (though via different pathways; see Heim, Ehlert, & Hellhammer, 2000; Raison & Miller, 2003; Miller et al., 2007; Sternberg, Chrousos, Wilder, & Gold, 1992). Evidence for the Role of Cortisol as a Biological Intermediary Evidence from a number of studies on stress-related changes in the output of cortisol has helped shape the notion that the HPA axis may serve as a biological intermediary linking early adversity to adult physical health. Meta-analytic findings of studies linking chronic stress to cortisol activity suggest that there is considerable variability (Miller et al., 2007), however. For example, connections between chronic stressors and HPA activity varied dramatically as a function of the timing of the stressor, whether the stressor involved trauma, whether the stressor was controllable or physically threatening, as well as individual emotional responses to the chronic stress (e.g., feelings of shame, loss). For example, individuals who faced physical stressors, such as war, and traumatic, uncontrollable stressors had high volumes of cortisol output over the day. In contrast, individuals facing chronic or controllable stressors did not differ from controls in their total daily output of cortisol. These findings suggest that cortisol is perhaps not best viewed as a generic marker of stress; it may, however, be a marker of stress under certain contexts and conditions (e.g., when experiencing traumatic and uncontrollable stressors). This view reinforces the findings of an influential meta-analytic study of cortisol responses to acute stress, which showed that levels of the hormone rise only in experimental settings that evoke uncontrollable social-evaluative threats (Dickerson & Kemeny, 2004). Notably, the best predictor of cortisol levels is the timing of a chronic stressor (Miller et al., 2007). Studies that focused on temporally distant stressors found lower-than-normal cortisol, whereas studies that focused on individuals' current experience of stress found elevations in cortisol levels. This finding suggests that stress might initially trigger a burst of cortisol, eventually leading to adaptations to the HPA axis that are meant to limit the damage that might result from sustained periods of heightened cortisol levels. In other words, it may be that the body mounts a counterregulatory response such that cortisol output rebounds below normal levels. This rebound effect makes sense in light of the fact that the HPA axis is regulated by a

potent negative feedback circuit, in which elevated levels of cortisol suppress the output of CRH and ACTH by acting on glucocorticoid receptors in the brain. Consistent with these meta-analytic findings, a number of studies have found that exposure to a wide range of early adversities, including socioeconomic disadvantage, neglect, abuse, and the loss of caregivers, is associated with later hypo-activation of the HPA axis. For example, in a prospective sample of adopted children, those who experienced severe neglect and abuse prior to the adoption (as reported by the adoptive parents) had lower morning cortisol values and flatter diurnal slopes in adulthood compared to individuals who did not experience neglect or abuse prior to the adoption (van der Vegt, van der Ende, Kirschbaum, Verhulst, & Tiemeier, 2009). Similarly, in a sample of older adults, Gerritson et al. (2010) found that the experience of any early adversity (e.g., abuse, poverty, parental divorce, loss) was associated with reduced morning cortisol levels and flatter diurnal slopes. Further, several studies have found evidence for a long-term effect of parental separation and loss on HPA axis functioning. In a small study of adult men, the loss of a parent in childhood was associated with greater cortisol levels throughout the day compared to men who did not lose a parent (Nicolson, 2004). Additional evidence suggests that prolonged separations from parents could be connected to adult HPA axis functioning (Kumari, Head, Bartley, Stansfeld, & Kivimaki, 2013). Using data from the British Whitehall II study, Kumari and colleagues (2013) found that the experience of prolonged maternal separation before the age of 16 was associated with flatter diurnal slopes compared to adults who did not experience prolonged maternal separation. Notably, these effects did not differ as a function of the reason for the separation (e.g., maternal death, wartime evacuation). Further, individuals who experienced parental separation exhibited larger cortisol awakening responses relative to individuals who did not experience parental separation. Some evidence suggests that socioeconomic disadvantage is associated with adult HPA axis functioning (e.g., Li, Power, Kelly, Kirschbaum, & Hertzman, 2007). For example, using data from the 1958 British cohort study, Li et al. (2007) examined participants' SES at birth and ages 7, 11, and 16. For men (but not women), low SES in childhood was associated with greater cortisol output as indexed by area under the curve, even when controlling for adult SES. Cortisol as a Biological Intermediary: Considerations and Caveats Although cortisol has regulatory influences on many bodily systems, these dynamics are highly complex. The result is that it is often difficult to interpret the biological significance of adversity-related differences in cortisol. Here we discuss these complexities in some detail, with the goal of facilitating more probative research on cortisol's role as an intermediary linking early adversity and later disease. We begin by noting that a number of steps must take place before the presence (or absence) of cortisol can begin to influence one of the body's tissues or organs. First, cortisol must bind to a specific receptor located inside a cell. Most cortisol that is released into the bloodstream, however, is attached to carrier proteins, such as cortisol binding globulin (CBG) or albumin.

When cortisol is attached to these proteins—which is about 90% of cortisol that is produced— it is unable to traverse cell membranes. Consequently, it cannot get inside the cell to bind to receptors. The other 10% of cortisol making its way through the bloodstream—which is free, or unattached to carrier proteins—must go through several steps before it is capable of modifying tissue. First, the hormone must traverse cell membranes and avoid being excreted by a structure called the multidrug resistance (MDR) pump, which expels molecules that threaten homeostasis. Further, it must avoid being converted to its inactive form by an enzyme called 11-β-hydroxysteroid dehydrogenase. Then it must bind to either a glucocorticoid or mineralocorticoid receptor, once those receptors shed some accessory molecules known as heat-shock proteins (see Raison & Miller, 2003, for a review of the mechanisms behind glucocorticoid signaling). Once cortisol has progressed through these steps, the newly formed receptor-hormone complex must translocate to the nucleus, where it is then capable of modifying the cell's program of gene expression. Clearly, several (often overlooked) steps must take place between the release of cortisol and the ultimate effect of cortisol on tissues of interest. An important implication of this multistep progression of cortisol activity is that even if a particular psychosocial factor of interest can be linked to the release of cortisol, the signal may not be heard downstream in the tissues of interest. One reason for this is that, as already discussed, most tissues are equipped with a host of counterregulatory mechanisms, such as carrier proteins (e.g., CBG) and the MDR pump that potentially can intervene in this process to ensure that acute changes in cortisol do not drastically disrupt homeostasis. This means that even if cortisol levels are increased markedly by a stressor, the tissues that are regulated by cortisol may not be affected because counterregulatory mechanisms may alter how loudly cortisol's signals are heard. For example, cells may alter the characteristics of the receptors to which cortisol binds. In addition, cells can reduce the density and activity of their receptors. Receptors that are downregulated or desensitized will pass on fewer signals to the nucleus of the cell, where gene expression occurs (Cole et al., 2007). Even in the absence of connections between a given stressor and cortisol, however, it is quite possible that there is ongoing activity at the level of the receptor that has important implications for physical health. In other words, the sheer volume of unbound cortisol in the body (and the amount that cortisol levels change in response to stressors) does not provide enough information to explain how the HPA axis could be involved in shaping physical health outcomes. One study nicely illustrates these complexities. In a sample of adult participants who varied in early-life SES (Miller et al., 2009), participants with low-status backgrounds had modestly elevated cortisol relative to high-status participants. Importantly, however, examination of gene expression profiles in leukocytes indicated that individuals from low-SES backgrounds had decreased glucocorticoid signaling relative to people with high early-life SES. In other words, the leukocytes of individuals from low-SES backgrounds were registering less cortisol signal. These findings are provocative because examination of cortisol levels alone would have suggested that low early-life SES confers risk for elevated circulating cortisol in adulthood, with the corresponding assumption that this extra cortisol could have a damaging effect on tissues and organs. Yet the gene expression findings suggested that individuals from low-SES

backgrounds had cells that were hearing less signal from cortisol, not more—a finding that indicates that examination of circulating levels of cortisol alone is an insufficient assessment of HPA axis activity. Thus, the study of cortisol in isolation—without also considering glucocorticoid signaling—does not provide an accurate picture of the ways in which HPA axis activity might serve as a biological intermediary linking early adversity and later health outcomes. An additional concern centers on the extent to which cortisol values exhibit within-person stability over time. Some recent studies highlighting cortisol's lack of temporal stability (Ross, Murphy, Adam, Chen, & Miller, 2014; Shirtcliff et al., 2012) call into question its plausibility as a mediator of the lengthy underlying pathologic processes that give rise to most chronic diseases of aging. For example, one recent study showed that only 13% of the variability in cortisol values over a 6-year period was attributable to a stable, traitlike component (e.g., Shirtcliff et al., 2012). Instead, over half of the variability in cortisol values could be explained by situational, or statelike, factors at the time of the assessment. Thus, even over relatively short periods of time, cortisol activity shows very little stability. In other words, if individuals' cortisol levels are not stable over relatively short intervals and appear instead to reflect dynamic changes in situational experiences of close temporal proximity, then it is unlikely that chronically dysregulated HPA axis activity is a mechanism that causally contributes to diseases such as cardiovascular disease, which evolves over decades. In summary, the HPA axis has been a popular target for mechanistic research on early adversity and adult disease over the last several decades. This explosion of research has led to a greater understanding of how responsive the system is to stressful experiences. Important features of cortisol, however, often have been overlooked in this research (e.g., glucocorticoid receptor availability and sensitivity, counterregulatory mechanisms that influence cortisol's actions in tissue, temporal stability and dynamics of cortisol). These features are important details that cannot be ignored, particularly if researchers want to develop a biologically plausible understanding of cortisol's role in explaining how early experiences influence chronic disease outcomes.

Allostatic Load Allostatic Load: Background and Measurement Some of our physiological systems, such as body temperature and blood pH, are tightly regulated, such that they operate within a narrowly specified range of functioning, with homeostatic set points around which the body attempts to achieve stability. Other systems, however, operate under much more flexible states, and they can change adaptively in response to the demands of the environment. In these circumstances, the systems operate under conditions of allostasis (“stability through change”). The main difference between allostatic systems and homeostatic systems is that allostatic systems have dynamic and adjustable mechanisms in place that are responsive to the environment, whereas homeostatic systems operate to achieve a specific set point despite the environmental demands. For example, during exercise, our bodies achieve thermoregulation through perspiration (thus achieving a fairly

static temperature despite changes in behavior), but our heart rate adjusts to the increased physical demands by beating faster. Another way of viewing allostasis is as a measure of plasticity. This ability to adapt to environmental demands enables temporary adjustments in physiological systems that allow for the ultimate goal of overcoming stressful or challenging situations. This plasticity in physiological systems, however, is thought to come at a cost: When these systems repeatedly are required to adapt to meet the demands of the environment—as in cases when individuals are exposed to chronic, repeated stressors—the systems may become inefficient or less flexible in their ability to adapt. The systems may not shut off completely, or there may be delays in the return to baseline functioning as a result of repeated environmental stressors. McEwen and colleagues (McEwen, 1998; McEwen & Seeman, 1999; McEwen & Stellar, 1993; Singer, Ryff, & Seeman, 2004) developed the concept of allostatic load, which refers to the wear and tear that results from repeated stressors that activate the body's regulatory controls in an attempt to achieve a new level of stability (and recovery from stress). Over time, this wear and tear manifests itself in the dysregulation of multiple physiological systems. Measures of allostatic load, therefore, sometimes are viewed as reflecting the number of physiological systems that are experiencing disruption in optimal functioning. A number of physiological systems are thought to be regulated under allostatic control, including the neuroendocrine, metabolic, cardiovascular, immune, and autonomic and sympathetic nervous systems (Singer et al., 2004). Accordingly, frequently a variety of biomarkers are included in measures of allostatic load. Neuroendocrine measures of allostatic load often include cortisol and catecholamines, such as epinephrine, norepinephrine, and dopamine. Immune measures often include markers of inflammation, including the cytokines interleukin-6 (IL-6) and tumor necrosis factor-α (TNF-α) as well as C-reactive protein (CRP), an acute phase protein that promotes inflammation by enhancing phagocytosis. Indicators of metabolic dysfunction include high-density lipoprotein cholesterol (this value is reversescored so that low values indicate risk), low-density lipoprotein cholesterol, triglycerides, glycosylated hemoglobin, and blood sugar. Cardiovascular components of allostatic load frequently include systolic and diastolic blood pressure and heart rate. Finally, waist-to-hip ratio or BMI usually are included in measures of allostatic load as indicators of adiposity. Researchers frequently create quartiles for each index, assigning individuals into dichotomous categories depending on whether their values are in the top high-risk quartile. These categories are then summed across indices to create a total composite score for allostatic load. One question related to this method of creating a composite score is whether the indicators of physiological dysregulation reflect a single underlying construct or if they are better conceptualized as separate components of physiological dysregulation. To answer this question, McCaffery, Marsland, Strohacker, Muldoon, and Manuck (2012) tested the factor structure of allostatic load indices. In this study, researchers found support for a single factor underlying the allostatic load indices, including measures of metabolic, inflammatory, and cardiac vagal activity measures. These findings lend support to the practice of using one composite score for the wide range of physiological systems that are included in assessments of allostatic load.

With that said, McEwen and others (e.g., McEwen & Seeman, 1999) have sorted allostatic load markers into categories, including primary mediators, secondary outcomes, and tertiary outcomes, in a way that is conceptually and methodologically useful. Primary mediators include chemical messengers, such as cortisol, dehydroepiandrosterone (DHEA), and catecholamines, which are thought to relay messages from the brain to organs and tissues that mobilize resources for coping with stress. Secondary outcomes are hypothesized to reflect the impact that repeated signaling from primary mediators has on target organs and tissues. These outcomes include tissue- and organ-specific indicators of health, including waist-to-hip ratio, blood pressure, glycosylated hemoglobin, and cholesterol. Finally, tertiary outcomes (which currently are not included in most measures of allostatic load) include actual diseases hypothesized to result in part from allostatic load (e.g., cancer, dementia, heart disease). The basic premise of this model is that primary mediators influence the development of secondary and finally tertiary (disease) outcomes. This hierarchical organization of allostatic load indicators offers a conceptually grounded approach to understanding how a stressor could contribute to wear and tear on the body and ultimately lead to the accumulation of chronic diseases. As far as we are aware, however, longitudinal studies have not yet been done to confirm the theorized progression from primary mediators to secondary and finally tertiary outcomes. Prospective studies of this sort will be useful for improving our understanding of the role of allostatic load in translating early adversity into disease. Evidence for the Role of Allostatic Load as a Biological Intermediary Allostatic load has been proposed as a biological intermediary linking early adversity to adult health (e.g., Repetti, Robles, & Reynolds, 2011; Taylor, Way, & Seeman, 2011). For example, in a prospective study of individuals in Sweden, Gustafsson, Janlert, Theorell, Westerlund, and Hammarstrom (2012) found that the experience of social stressors (e.g., loss of a parent, social isolation) and material adversity (e.g., parental unemployment, poor standard of living) in adolescence and early adulthood were uniquely predictive of allostatic load in midlife. Similarly, using data from the MIDUS study, Gruenewald et al. (2012) found that low SES in childhood and adulthood predicted allostatic load in adulthood. Notably, low SES at each developmental period was associated uniquely with later allostatic load. Further, Gruenewald and colleagues found evidence of a cumulative effect: Individuals who were persistently low in SES across childhood and adulthood had greater allostatic load relative to individuals who were consistently high in SES or who were upwardly mobile. Interestingly, these effects remained significant even after controlling for health behaviors and psychosocial variables in adulthood. Evidence from studies with children and adolescents suggests that these processes might start early in life (e.g., Evans & Kim, 2012). Allostatic load, in turn, has been shown to be predictive of cardiovascular disease and mortality (Seeman, McEwen, Rowe, & Singer, 2001; Seeman, Singer, Rowe, Horwitz, & McEwen, 1997). These links are not particularly surprising, given that the index includes measures of blood pressure and central adiposity, two of the most established risk factors for cardiovascular disease and health. Reductions in allostatic load have been associated with lower 7-year mortality risk (Karlamangla, Singer, & Seeman, 2006). These findings have led

researchers to suggest that allostatic load may be a pathway through which exposure to early adversities can lead to greater health problems in adulthood (e.g., Danese & McEwen, 2012; Gruenewald et al., 2012). Indeed, one strength of the allostatic load model is that the composite score reflects the extent to which multiple physiological systems are operating under strained conditions. If early adversity results in different biological cascades across individuals, leading to a wide variety of endpoints with possible dysregulation, then a single biological marker might not always be linked to adverse experiences. By creating a composite measure of allostatic load, researchers are better positioned to capture the connection to early adversity. Thus, in a way, the practice of creating a composite measure of allostatic load provides a simplified way of testing for multifinality in the links between early adversity and long-term health outcomes by creating a single outcome variable that reflects a diverse array of physiological systems. Allostatic Load as a Biological Intermediary: Considerations and Caveats One of the concerns about allostatic load, as it has typically been operationalized in this literature, is that its components vary in how proximal they are to actual disease. As McEwen and Seeman (1999) emphasize, primary mediators, such as cortisol, probably are quite distal causes of physical health problems, which act through more proximal secondary outcomes, such as insulin resistance, to cause disease. Yet most studies to date group all of these components into a single aggregate, making the implicit assumption that they are equally powerful and equally detrimental in terms of long-term health. We recognize that there is factor-analytic evidence to support a single composite score for allostatic load (McCaffery et al., 2012). However, it still seems likely that allostatic load components vary considerably in how proximal, necessary, and sufficient they are in the disease process. Indeed, allostatic load might foreshadow cardiovascular disease not because individuals are elevated on a diverse set of physiological measures but because they have high blood pressure and excess central adiposity, which are well-established risk factors for this condition. This lack of clarity in the utility of allostatic load indices makes it difficult to know whether this measure provides incremental value over already established stand-alone risk factors for poor health and whether some components may be more powerful contributors than others for poor health outcomes. Moreover, the practice of creating nonstandardized cut points for dichotomizing the indicators raises several methodological issues. For one, this practice often is done using sample-specific cutoffs, such as top quartiles, which makes it difficult to compare across studies and increases the odds of chance findings. Another problem with the cutoff approach is that it ignores the fact that functioning and risk within each domain may follow a more linear gradient. If domains have validated cutoffs with prognostic significance (e.g., blood pressure, cholesterol), then these established criteria may be more meaningful indices to use when sorting individuals into high- and low-risk categories. One final concern relates to the choice of measures included in allostatic load composites. As described earlier, indices of allostatic load are categorized into primary mediators, secondary outcomes, and tertiary outcomes. Yet this level of categorization is not reflected in the single composite measure of allostatic load. Instead, these markers are aggregated across the different

stages without consideration of the relative importance of each indicator or acknowledgment that primary mediators should prospectively predict secondary outcomes, which should in turn shape the onset and progression of diseases. Given that the hierarchy is so fundamental to the theory behind allostatic load, it is unclear why these indicators are treated equally. In fact, some measures in and of themselves may pose no clear health risk, or do so in only a very distal manner, whereas other indices (e.g., blood pressure) are actually measures of damage already done to the system. For example, having higher-than-average levels of dopamine is not considered a risk factor on its own, but elevated blood pressure and cholesterol are risk factors for cardiovascular disease. Based on the theoretical model, early adversity should be related most closely to the primary mediators of allostatic load, and these changes should presage the onset of secondary outcomes. This kind of cascading model has not been tested to date, however. In summary, allostatic load has been implicated as a potential biological mechanism linking early adversity to later disease and early mortality. We look forward to the next generation of research on this construct, which we hope will clarify some of the complex issues surrounding conceptualization and measurement of allostatic load. Future studies that can clarify the clinical significance of allostatic load and the processes through which allostatic load contributes to disease will be especially useful.

Telomeres Telomeres: Background and Measurement Recently there has been an active interest in the ways in which psychosocial stressors might accelerate cellular aging through changes in telomere structure (Aubert & Lansdorp, 2008; Aviv, 2008; Mather, Jorm, Parslow, & Christensen, 2011; Price, Kao, Burgers, Carpenter, & Tyrka, 2013). Telomeres are nucleoproteins located at the end of chromosomes. Telomeres act like caps that prevent DNA degradation when cells divide and in doing so ensure the stability of chromosomes and the cells themselves. Briefly, when cells divide, replication enzymes are unable to function at the very end of chromosomes (an issue known as the end replication problem; De Lange, 2009). Without telomeres, each cell division would result in segments of chromosomal DNA being lost to replication, meaning that as time went on, progressively more of genetic code (DNA) would disappear. The presence of telomeres slows the progressive erosion of genetic material. Telomeres serve as a kind of buffer for the chromosome, which can absorb successive losses of material with each replication. Because telomeres are comprised of noncoding DNA, or material that is not essential for building proteins, their degradation does not have significant consequences for cellular function in the short term. Moreover, as telomeres shorten with successive replication cycles, cells produce an enzyme, known as telomerase, which acts to restore their length. However, most cells do not contain enough telomerase to maintain telomere length indefinitely. So, with time, telomeres degrade. When this occurs, healthy cells enter a phase known as senescence, where they cease replication and can have limited functional capacity. Thus, cellular senescence is the endpoint of a lengthy process involving repeated cell replications and waning telomerase activity to

counterbalance it. Evidence for the Role of Telomere Degradation as a Biological Intermediary Telomere length has been viewed by some as a biological marker of aging (Aubert & Lansdorp, 2008; Epel, 2009; Mather et al., 2011). Indeed, health-compromising behaviors, such as smoking, have been related to shorter leukocyte telomere lengths. And to the extent people have shorter leukocyte telomeres, their risks for morbidity and mortality from various conditions increase (e.g., Kimura et al., 2008; Valdes et al., 2005). Some researchers argue that shorter telomeres (relative to telomeres of other individuals of the same age) are an indication of accelerated aging, with the potential for long-term health consequences. Emerging evidence suggests that early exposure to adversity may be associated with telomere degradation (Drury et al., 2012; Kananen et al., 2010; Tyrka et al., 2010). For example, Tyrka et al. (2010) found that recalled emotional and physical neglect in childhood were associated with shorter telomeres in white blood cells. Similarly, Kananen et al. (2010) reported a connection between early adversity (e.g., parental unemployment, parental mental health problems) and telomere length in white blood cells in a sample of anxious and nonanxious adults. Further, in a sample of Romanian children, Drury and colleagues (2012) found a link between buccal cell telomere length in middle childhood and time spent in early institutional care. Additional evidence for a link between early exposure to adversity and later telomere attrition comes from Kiecolt-Glaser and colleagues (2011), who found that older adults who had experienced multiple childhood adversities had shorter telomeres in white blood cells compared to adults who were not exposed to early adversities in childhood. Some evidence suggests that supportive parenting experiences may play a role in buffering children from telomere degradation associated with early adversity. In a cross-sectional study, Asok, Bernard, Roth, Rosen, and Dozier (2013) found that children from high-risk families who experienced responsive parenting had greater buccal cell telomere length than children from high-risk families whose parents were less responsive. These findings offer some insight into the ways in which telomere length may be influenced by both supportive and unsupportive environmental contexts, although longitudinal studies will be needed to lend support to the notion that these environmental inputs play a causal role in affecting telomere length. Telomere Degradation as a Biological Intermediary: Considerations and Caveats The studies just reviewed consistently find linkages between early adversity and telomere length, which suggests the provocative hypothesis that accelerated cell aging might be a mechanistic pathway to later disease. However, as is the case with the broader literature on telomeres, several features of these studies complicate interpretation. First, all of the studies were based on a single measurement of telomere length. There are significant individual differences in telomere length across the population (Aviv, Valdes, & Spector, 2006; Geronimus et al., 2010; Takubo et al., 2002). Thus, with a single time point measure, we cannot be certain whether the disparities in telomere length are a consequence of the earlier adversity or are simply a correlate. For example, it could be that the same children who are

exposed to adversity are born with relatively short telomeres, for partly or fully unrelated reasons. To sort this complexity out, multiwave studies or natural experiments are needed. Along these lines, some of the most convincing evidence that exposure to adverse conditions in childhood accelerates telomere degradation comes from the Environmental-Risk Longitudinal Twin Study, a nationally representative sample of children in Great Britain (Shalev et al., 2013). Researchers assessed children's exposure to violence in childhood, including domestic violence, bullying and victimization, and maltreatment from an adult. Telomere length in buccal cells was assessed at baseline (age 5) and 5 years later (age 10). Shalev and colleagues (2013) found that children who were exposed to two or more types of violence had accelerated telomere erosion across the 5 years compared to children who were exposed to one type of violence or who were not exposed to violence. These are provocative findings from a methodologically rigorous study. They represent an important starting point for mechanistic research on early adversity. Nevertheless, there is some uncertainty about the clinical relevance of telomeres in buccal cells for the kinds of chronic diseases of aging we are concerned with in this chapter. For those conditions, telomere length at the site of disease pathogenesis—for example, in the heart for cardiovascular disease, the pancreas for diabetes, the immune system for virally mediated cancers and infectious diseases—would be more germane. It is possible that buccal cells, or other easily accessed specimens, such as white blood cells, can serve as a proxy for telomeres in these tissues of interest. But given the degree of tissue specificity seen in most cellular processes, this seems like a dubious assumption to proceed on, at least until the field has accumulated sufficient evidence to justify it. Indeed, recent studies indicate that the correlations between telomere length in buccal and immune cells are modest at best (Gadalla, Cawthon, Giri, Alter, & Savage, 2010). This problem is not limited to buccal cells. As an alternative, researchers often measure telomere length in leukocytes, based on the quite reasonable presumption that immune cell aging could have deleterious consequences for host resistance to infections and cancers. But even when leukocytes are collected, most studies assess telomere length in bulk populations with heterogeneous cellular content. Peripheral blood contains a number of different types of leukocytes, including granulocytes, lymphocytes, and monocytes. These cell types differ significantly in their telomere profiles (Aubert & Lansdorp, 2008). What makes this methodological decision even more problematic is that people vary in the relative proportion of these cell types in peripheral blood, and adversity is known to shift that balance (Segerstrom & Miller, 2004). Thus, any linkage between adversity and telomeres could reflect individual differences in the balance of cell types present in the assay rather than cellular aging per se. To rule out this possibility, researchers need to account for cellular heterogeneity through statistical controls, or, even better, physically sort cells into different phenotypic clusters prior to telomere analysis (e.g., see research on stress and telomeres by Cohen et al., 2013; Damjanovic et al., 2007). Aside from the heterogeneity in cell types used for telomere assessment, it is important to consider the fact that not all cells in the body replicate. Of the major immune cells, only

lymphocytes do so routinely. This fact raises the question of what telomere length means in samples that are mainly composed of nonreplicating cells. (This is the case with most whole blood analyses of telomeres, where roughly 50% to 80% of the cells are nondividing granulocytes or monocytes.) Some researchers have argued that in nondividing cells, such as granulocytes, telomere length might index senescence of the parent cell—the cell that divided to form the granulocyte. If so, these measures would take on significant new meaning. But until this issue and others are settled, we remain cautious about the interpretation of the telomere data.

Epigenetics Epigenetics: Background and Measurement All cells within an individual carry an identical DNA sequence, which is established at conception and fixed for life. (Lymphocytes and cells that have acquired mutations are exceptions to this rule.) The DNA sequence serves as a blueprint for transcription, the process whereby cells synthesize ribonucleic acid (RNA) molecules—a process known as gene expression. RNA molecules are later translated into proteins, which cells use for structural and functional purposes. Epigenetics describes stable changes in gene expression activity that arise without changes to the gene's DNA sequence (Jaenisch & Bird, 2003; Jirtle & Skinner, 2007). A primary function of epigenetic alterations is to allow cells to develop and maintain specialized functions. For example, epigenetic alterations can modify a cell's ability to transcribe a particular gene into RNA. Because RNA serves as a template for translation of proteins, these epigenetic alterations have downstream influences on how much of the gene's protein ultimately is synthesized. When this process takes place across many different genes, it can give rise to significant phenotypic diversity in cells. This variability is thought to play a role in the long-term development of physical disease (e.g., Rakyan, Down, Balding, & Beck, 2011), although the nature of these pathways is still largely unknown. Epigenetic modifications to DNA typically occur in one of two ways (Whitelaw & Garrick, 2006). The first modification is DNA methylation, which involves the attachment or removal of a methyl group to cytosine residues in a gene's promoter. The methyl groups prevent transcription factors from interacting with DNA to modulate gene expression, which makes the gene inactive. The second form of epigenetic modification involves changes to the chromatin structure that packages the DNA. This process occurs by attaching or removing chemicals from the histone proteins that hold DNA within the cell's nucleus. These proteins cause the DNA near the gene to become more or less tightly coiled, which makes it more or less difficult for RNA polymerase and transcription factors to access its promoter (Whitelaw & Garrick, 2006). These epigenetic modifications can be quite stable over the life span, and sometimes they are passed along to daughter cells during mitosis. As such, it is possible for early experiences— such as early exposure to adversity—to create lasting epigenetic modifications that persist over the life span, which may help explain how early experiences become embedded in the epigenome and contribute to chronic diseases many decades later. Measurement of epigenetic modifications typically relies on DNA microarray analysis. This process assesses the presence

of methylation at multiple promoter sites in each of many thousands of different genes (e.g., these analyses can quantify the proportion of sites that are methylated).

Evidence for the Role of Epigenetic Modifications as a Biological Intermediary Recent evidence suggests that epigenetic processes might serve as pathways through which early experiences can bring about long-term changes in the activity of genes, thereby contributing to the pathogenesis of disease (e.g., Roth, Lubin, Funk, & Sweatt, 2009). Much of the early work in this area centered on animal models, including the work by Meaney and colleagues (see Meaney, 2001, for a review). Meaney found that, in rats, certain caregiving experiences early in infancy were associated with long-term changes in rat pups' stress response physiology (Liu et al., 1997). Specifically, rat pups that received high levels of maternal licking and grooming and arched-back nursing showed attenuated cortisol responses to threat and increased exploratory behavior—effects that lasted into adulthood (Caldji et al., 1998; Liu et al., 1997). In a series of elegant mechanistic studies, Meaney and his colleagues discovered that these lasting differences in stress responsivity arose through epigenetic changes to glucocorticoid receptor genes in the hippocampus of the offspring (Weaver et al., 2004)—in other words, in the rat pups that received more licking and grooming. In essence, maternal nurturance in the early days of life led to epigenetic changes in the offspring's hippocampus including demethylation of DNA and acetylation of histone proteins, which facilitated expression of the glucocorticoid receptor gene across the life span (Weaver et al., 2004). Greater expression of this receptor, in turn, enabled tighter regulation of the HPA axis, the hormonal system that controls the release of cortisol in response to stress. Notably, Weaver et al.'s (2004) findings revealed that the epigenetic modifications that took place during the first week of the rat pups' lives persisted into adulthood, suggesting considerable stability in these early programming effects. Recent work with humans suggests that early experiences—including adversity—also bring about epigenetic changes that affect gene expression (e.g., Bick et al., 2012; McGowan et al., 2009; Oberlander et al., 2008). In one study, hippocampal tissue from suicide victims was assessed postmortem (McGowan et al., 2009). Epigenetic modifications varied as a function of whether participants had been abused as children: Individuals who had been abused had greater methylation at sites in the glucocorticoid receptor gene promoter. In a prospective study of maternal depression and infant cortisol reactivity, Oberlander and colleagues (2008) found that exposure to prenatal maternal depression led to epigenetic changes in infants' gene expression that were associated with elevated stress reactivity in the HPA axis. Specifically, prenatal exposure to depression during the third trimester was associated with increased methylation at sites in the glucocorticoid receptor gene promoter. Further, methylation at this location was related to heightened cortisol reactivity when the infants were 3 months old, suggesting that this epigenetic alteration may have functional implications for HPA axis regulation. Additional evidence for lasting epigenetic modifications as a result of early adversity comes from a sample of adults who were exposed to famine prenatally as a result of wartime food scarcity and their unexposed same-sex siblings (Heijmans et al., 2008).

Compared to their unaffected siblings, prenatal exposure to famine was associated with reduced DNA methylation in the promoter of the insulin-like growth factor II gene, which plays a role in insulin- and growth-related activities. Additional evidence from studies of early adversity and epigenetic modifications also supports the notion that adverse experiences are biologically embedded in the epigenome (Borghol et al., 2012; Essex et al., 2013; Tyrka, Price, Marsit, Walters, & Carpenter, 2012). For example, Essex and colleagues studied DNA methylation in the buccal cells of 109 adolescents from a larger longitudinal study. In this sample, the extent to which adolescents' buccal cells were methylated varied as a function of their parents' stressors years earlier. The timing of stress influences differed for mothers and fathers. Specifically, methylation during adolescence was related to maternal stress during infancy but paternal stress during preschool. These findings are consistent with those of Tyrka and colleagues (2012), who found connections between early-life familial adversity (e.g., parental death, maltreatment) and methylation of the glucocorticoid gene promoter in white blood cells. Adults who, as children, experienced early familial adversity had increased methylation of the promoter region relative to adults who experienced more stable and nurturing care in childhood. Borghol and colleagues (2012) similarly identified connections between early SES and methylation of white blood cells in 40 participants from the 1958 British cohort study. In this detailed microarray analysis of approximately 20,000 genes, over 1,200 gene promoters were differentially methylated as a function of early-life SES—over twice as many regions that were differentially methylated as a function of adult SES. Lam et al. (2012) posed similar questions about early SES in a sample of healthy adults aged 25 to 45 but found a more modest association with methylation patterns. These findings suggest that early adversity may leave an epigenetic residue in the genome of some cells. Based on basic research in model systems, there is reason to suspect these residues may be durable and persist over lengthy periods of time. This stability could help explain how early experiences can develop into diseases decades after the exposure to adversity has subsided. With that said, we do not yet have the prospective, longitudinal evidence in humans to know whether the findings reflect this kind of lasting biological imprint versus some other process or methodological artifact.

Epigenetic Modifications as a Biological Intermediary: Considerations and Caveats Like the other biological mechanisms we have discussed, the study of epigenetic modifications to DNA poses some important methodological challenges that will need to be addressed in future studies. For example, in trying to extend Meaney and colleagues' findings (Weaver et al., 2004) to humans, research has made use of cord blood cells, white blood cells, and buccal cells. These approaches make sense logistically. Hippocampal tissue cannot be noninvasively acquired in living humans, whereas cells from the blood, mouth, and umbilical cord can be extracted noninvasively. With that said, it is important to keep in mind that very little is known about how well these tissues can serve as a proxy for the extent of methylation in the hippocampus. Many researchers believe the epigenetic processes exist mainly to allow tissues

to functionally specialize. If that is true, we need to be cautious about assuming cross-tissue consistency in epigenetic marks, at least until there is evidence to support doing so. Further, the use of heterogeneous pools of cells for measuring DNA methylation raises some concerns. As was the case with telomeres, people vary in the relative proportion of cell types that make up their immune systems and their cord blood, and stress can shift the balance of cell types. And as was the case with telomeres, the degree of methylation varies across cell types (Lam et al., 2012; Ohgane, Yagi, & Shiota, 2008). Together, these issues complicate interpretation of the existing epigenetic data in humans—any disparities linked to adversity could reflect differences in the cellular composition of samples or the degree of methylation. Some studies have addressed the issue of cellular composition in their analyses. Lam and colleagues (2012), for example, found that white blood cell methylation was correlated with the relative amount of lymphocytes and monocytes present in individuals. Thus, assessments of DNA methylation may need to take into account the relative proportion of cell types in the sample. Finally, it will be important for future studies to explore the functional implications of adversity-related changes in methylation. Much work in this area proceeds on the assumption that epigenetic changes will have consequences for gene expression, as they do in the animal studies. But this assumption is not always borne out in studies of humans, which rely on large heterogeneous pools of cells from the buccal cavity or the immune system. For example, in the study by Lam et al. (2012), associations between methylation and transcriptional profiles were very modest, and this has been the case in other studies too. These findings suggest that methylation is likely to be one process among many that are important in regulating gene expression patterns. Thus, future epigenetic research in this area would be most helpful if it supplemented methylation data with studies of the functional consequences of methylation for gene expression or other relevant outcomes (e.g., see Oberlander et al., 2008, study). Also important in future research will be consideration of other epigenetic mechanisms—such as histone modifications and microRNA expression—that can work in concert with methylation to influence patterns of gene expression.

Inflammation Inflammation: Background and Measurement Inflammation occurs when cells of the innate immune system, including neutrophils, dendritic cells, monocytes, and macrophages, gather at the site of an infection or injury. These cells attempt to eliminate the pathogen, rid the body of infected tissue, repair any damage the pathogen caused, and begin the process of healing. The inflammatory response is essential for survival. Without it, minor injuries or infections would be lethal. However, the inflammatory response must be regulated carefully; otherwise, inflammation can become persistent and contribute to the emergence of multiple diseases of aging. Inflammation is orchestrated by signaling molecules known as cytokines, which are released by immune cells and the damaged tissue. The major cytokines involved with inflammation are interleukin (IL)-1, IL-6, and TNF-α. Researchers sometimes use concentrations of these molecules in circulation as a rough estimate of ongoing inflammatory activity. However, these

cytokines are fairly unstable in blood, so a more common approach is to measure CRP, a molecule produced by the liver during inflammation. CRP provides a reliable index of lowgrade chronic inflammation over the preceding month or so and is prognostic of morbidity and mortality from a number of chronic diseases of aging, such as diabetes, obesity-related problems, cardiovascular disease, autoimmune disease, and cancer (Danesh, Collins, Appleby, & Peto, 2000; Libby, Ridker, & Hansson, 2009; Ridker, 2007; Yeh & Willerson, 2003). CRP's role is particularly well established in the progression of cardiovascular disease, where in apparently healthy individuals, it presages disease risk in a roughly dose-response manner. The role of inflammation in such a broad array of diseases, along with the relative ease of assessment, has prompted researchers to consider it as a mechanism linking early adversity to later disease. Evidence for the Role of Inflammation as a Biological Intermediary Mounting evidence supports the idea that exposure to early adversities is linked to measures of chronic inflammation. For example, in a prospective study of 12,000 adults from diverse backgrounds, early-life SES measured in childhood was inversely associated with CRP, independent of current SES (Pollitt et al., 2007). Similarly, in the British Women's Heart Health Study, evidence for a cumulative effect of SES was found: Lower status across development was associated with increases in CRP (Lawlor, Smith, Rumley, Lowe, & Ebrahim, 2005). Further, links between childhood SES and inflammation were examined in a sample of adults who took part in the Coronary Artery Risk Development in Young Adults study (CARDIA; Taylor, Lehman, Kiefe, & Seeman, 2006). In this study, Taylor, Lehman, Kiefe, and Seeman (2006) found that childhood SES was associated prospectively with measures of psychosocial functioning (e.g., depression, social contacts), which in turn predicted CRP. Additional evidence suggests that early-life socioeconomic disadvantage confers risk for immune system profiles characterized by proinflammatory responses (Miller et al., 2009). In this study, researchers examined healthy adults who were either low or high in childhood SES (as defined by the prestige of their parents' occupations). Participants' white blood cells were cultured in vitro with a series of microbial products, and the magnitude of the cell responses to these stimuli was indexed by subsequent production of IL-6. Participants who were exposed to socioeconomic disadvantage as children showed more pronounced IL-6 production in response to bacterial challenges relative to participants from families of high SES. Notably, participant groups in this study were balanced on the prestige of their adult occupations, which removed the possibility that these findings were attributable to adult SES. Similar evidence has been found for the role of the early family emotional climate on adult inflammation. In the Dunedin longitudinal study, for instance, multiple assessments of early maltreatment and harsh parenting were predictive of CRP at age 32 (Danese et al., 2009; Danese, Pariante, Caspi, Taylor, & Poulton, 2007). For example, Danese and colleagues (2007) reported that adults who had been maltreated as children had a threefold higher level of CRP at age 32 relative to individuals who had not been maltreated. Similarly, those who came from poor backgrounds and were socially isolated as children were also more likely to show

high CRP; notably, these risk factors were independent of each other. Several other studies have assessed inflammation in samples of adults who were asked to report retrospectively on experiences of childhood abuse. In one study, adults who had been abused as children had higher circulating IL-6 and TNF-α levels relative to adults who did not report being abused as children (Kiecolt-Glaser et al., 2011). In the Nurses' Health Study II, Bertone-Johnson and colleagues found a link between retrospectively reported sexual abuse in adolescence and CRP and IL-6 in adulthood (Bertone-Johnson, Whitcomb, Missmer, Karlson, & Rich-Edwards, 2012). Interestingly, however, this pattern did not emerge when examining the link between inflammation and reports of physical abuse. In the MIDUS sample, Slopen and colleagues (2010) examined a range of adverse early-life experiences as predictors of inflammation. In particular, adversity in childhood and adolescence was measured across three domains, including stressful life events (e.g., parental alcohol or drug problems, school failure), parent-child relationship quality (including relationships with mothers and fathers), and the frequency of verbal and physical abuse. Scores across each of these domains were standardized and then summed to create a composite of early-life adversity. This composite measure was associated with five different biomarkers of inflammation for African-American participants but not for participants of European descent (Slopen et al., 2010). In summary, emerging evidence suggests that to the extent that individuals are exposed to early adversity, as adults they display evidence of mild, chronic inflammation. Further, many of the chronic diseases of aging (e.g., cardiovascular disease, strokes, autoimmune diseases, some forms of cancer) are widely recognized to involve chronic inflammation. Thus, inflammation could play a mediating role in the links between early adversity and adult physical health. Inflammation as a Biological Intermediary: Considerations and Caveats Converging evidence suggests that early adversity may contribute to later disease through inflammatory processes. Despite these encouraging findings, we note that there are several limitations that will need to be addressed in future studies. Like the studies of other biological mechanisms, many of the studies linking early adversity to inflammation rely on a single assessment of inflammation, which limits our ability to conclude that exposure to adversity causally shapes inflammatory activity. Longitudinal studies, with multiple assessments of inflammation, will help clarify the extent to which early stressors lead to changes in inflammation over time. These types of studies also will help to clarify the chronicity of the inflammation. To be plausibly involved in accelerating disease, the inflammation would need to be long standing. Only via multiwave studies can we ascertain whether inflammation is transitory or chronic. Another issue to be addressed in future studies is the fact that most existing studies of inflammation capture ongoing inflammatory activity in peripheral blood rather than inflammation in tissues or organs. Thus, it is unclear whether these measures of inflammation reflect activity happening in the tissues that are directly relevant for chronic diseases (e.g., coronary arteries, pancreas). Finally, nearly all studies focus on proinflammatory processes, but anti-inflammatory signals play an important role in regulating

the balance of inflammatory activity. Going forward, researchers should examine both pro- and anti-inflammatory signals to capture a more comprehensive picture of how the immune system responds to adversity.

Concluding Comments and Directions for Future Research Over the last 30 years, researchers have made considerable progress in identifying links between early adversity and adult physical health outcomes. Findings from epidemiological studies, animal studies, and naturalistic experiments converge on the idea that exposure to early adversity is a risk factor for cardiovascular disease, metabolic disruptions, some cancers, and even early mortality. Great strides have been made in formulating hypothesisdriven, conceptually based models that set out to test the processes by which early stressful experiences become embedded in the body only to emerge as physical health problems, early aging, and chronic disease several decades later. Yet despite this wave of exciting research progress, we are only beginning to understand the pathways linking early adversity and health. Many of the studies to date, including retrospective reports and cross-sectional study designs, have significant methodological limitations, which limit our ability to draw strong conclusions about causality. Similarly, our knowledge about the biological mechanisms that might explain the progression from early adversity to accelerated aging and disease in adulthood is considerably lacking. Next we identify several key research areas that will be important to explore in the next generation of studies linking early adversity to adult physical health. The list of research priorities is far from exhaustive; nevertheless, advancements in each of these areas will increase our understanding of how and under what contexts early adversity shapes long-term physical health outcomes.

Expanding Research to Other Periods of Development The large majority of the research reviewed in this chapter focuses on the impact of adversity that occurs during the early childhood years. There is good reason for this emphasis: During these early years, many biological systems are especially sensitive to environmental effects, and many organ systems and tissues are undergoing a rapid period of development, expansion, and specialization. At the same time, however, these early years capture only a brief glimpse of development, and we believe that other developmental periods as well as transitions across developmental periods may be particularly sensitive to the effects of adversity in ways that can make individuals susceptible to physiological disruptions and progression toward disease. We describe several of these opportunities for future research next. Prenatal Period The fetal-programming model (Lucas, Fewtrell, & Cole, 1999) posits that in utero experiences shape infants' development by exposing them to maternal signals about

environmental conditions to be expected at the time of birth. From an evolutionary perspective, the ability to send signals to the developing fetus has a number of advantages for survival: The fetus, in response, can alter its development in preparation for the demands of the environment. Cottrell and Seckl (2009) have suggested that prenatal overexposure to glucocorticoids might be one mechanism through which prenatal stress results in greater risk for health problems, such as cardiovascular disease and metabolic abnormalities. Evidence from animal studies lends support to this hypothesis, but whether such an effect is present in humans remains to be seen. Findings from a number of studies suggest an important link between prenatal experiences and subsequent outcomes in infants (e.g., Field et al., 2004; Oberlander et al., 2008; Sandman, Davis, & Glynn, 2012; Sharp et al., 2012) and adults (Heijmans et al., 2008). Using a sample of mothers who varied in their levels of depression during and after pregnancy, Sandman and colleagues (2012) examined the hypothesis that infants are calibrated in utero for postnatal environmental conditions. As predicted, infants with the best performance on measures of mental development (e.g., sensory-perceptual acuity) and psychomotor development (e.g., body control) were ones who had similar pre- and postnatal environmental experiences—even when those experiences included maternal depression. In other words, infants who experienced maternal depression pre- and postnatally fared better than infants whose mothers were depressed at only one time point. These somewhat counterintuitive findings suggest that infants receive important programming signals in utero, and when these signals do not correspond with the external environment, infants show early delays in motor skills and mental development. Evidence about the long-term deficits related to the incongruence between prenatal and postnatal environments is lacking, however. It is possible that these effects disappear by childhood or have little long-term health relevance. Nevertheless, continued research in this area will be useful for determining the extent to which prenatal exposure to adversity confers risk for long-term physical health problems. Adolescence Adolescence is a developmental period of multiple transitions, characterized by fundamental changes in biological, cognitive, social, and psychological processes (Holmbeck, 1994; Lerner & Castellino, 2002). Researchers now widely consider adolescence to be the second critical period of development, and experiences during this period are thought to have an unusually strong influence on long-term health outcomes. This plasticity confers unique sensitivity to environmental and contextual experiences that can have a lasting impact on health. Indeed, from a health perspective, adolescence represents a critical time in which long-term health trajectories are set in motion, and the precursors of chronic disease often emerge during this period (Lule, Rosen, Singh, Knowles, & Behrman, 2006). Further, recent estimates suggest that a growing number of adolescents show symptoms that are indicative of metabolic disruption (Cornier et al., 2008). Compared to the extensive body of research on pediatric and adult health, adolescent health has received far less attention, which is surprising given that adolescents represent 25% of the world's population (Sawyer et al., 2012). The World Health Organization, UNICEF, and the

Lancet (Sawyer et al., 2012; United Nations Children's Fund, 2012; World Health Organization, 2014) recently highlighted the importance of investing in adolescent health research in order to address significant gaps in knowledge about determinants of health across the life span. The growing prevalence of health problems in childhood and early adulthood, including obesity and diabetes, has both immediate and long-term social, economic, and medical consequences. Although adolescent mortality rates are quite low, the precursors of chronic disease often emerge during this period (Balagopal et al., 2011; Berenson & Srinivasan, 2005). For example, risk for heart disease, including the early formation of atherosclerotic plaques, becomes apparent in adolescence. Thus, attempts to identify the processes that set adolescents on trajectories that culminate in health problems across the life span are of significant public health and economic importance. Moreover, identification of early-onset predisease processes has the potential to identify individuals for whom targeted interventions may be most beneficial. An additional question about the connection between stressors and health outcomes in adolescence relates to the fact that the onset of puberty corresponds with many of the transitions that take place during this developmental period. We know little about how puberty and the corresponding shifts in stress hormones interact with stressful experiences to shape physical health outcomes—both in adolescence and into adulthood. These questions will be important to address in future research. The Study of Transitional Periods Across Development Transitional periods in development, such as the transition from elementary school to middle school and the transition to adolescence, have the potential to be stressful periods for children (Graber & Brooks-Gunn, 1996). These periods often are marked by dramatic shifts in behavior, responsibility, rules, and opportunities. Given the sweeping changes across social, cognitive, affective, and biological domains and the onset of new stressors that take place around developmental transitions, researchers should examine how transitions may create stressors that put children at risk for health problems. Transitional periods that affect social and emotional functioning also may engender changes in physiology and risk for poor health. To study biological changes that co-occur with transitional periods, longitudinal studies will be necessary, an issue we discuss in more detail later.

Strengthening Study Designs Throughout this chapter, we have noted some of the important methodological limitations that limit our ability to make definitive conclusions about the extent to which early exposure to adversity leads to chronic diseases of aging in adulthood. Next we review a few of the main study design limitations and offer some solutions for stronger research designs that can be used to test research questions about how early adversity becomes biologically embedded in the body and translates into physical disease in adulthood. To date, cross-sectional and retrospective studies make up the bulk of studies on early adversity, biological intermediaries, and later health outcomes. Although there certainly is

merit in identifying connections at the cross-sectional level, our ability to draw conclusions about the causal nature of these links is substantially limited with these study designs. For one thing, retrospective reports have significant weaknesses (as discussed earlier in the chapter), which call into question the validity of the reports. Further, cross-sectional study designs cannot shed light on whether exposure to adversity contributes to dysregulation and disease or is simply a correlate of these outcomes. Future research needs to incorporate longitudinal study designs with multiple assessments of biology to allow for stronger statements about the direction of effects. These study designs can be costly and time-intensive, but they are important for researchers who wish to take a developmental approach to understanding the processes through which early adversity influences health. Of course, it is not necessary for researchers to conduct decades-long research investigations focused on experiences from infancy through adulthood in order to add to our understanding of adversity and health. For researchers who wish to test mediational models and developmental cascades models, at least three time points are preferable, with sufficient time between assessments for measures of interest to change during that period. Longitudinal designs that are conducted over relatively short time frames still can add insight into the intermediary processes that take place on the road from early adversity to chronic disease. Another concern focuses on the nature of various proposed biological mediators of the link between early adversity and adult health. Many of the biological mechanisms that we have described in this chapter are thought to serve as intermediate phenotypes, and they should be assessed at multiple time points to examine change over time. Research that can map the trajectories of these biological markers across time would greatly add to our understanding of how these mechanisms eventually lead to disease. Moreover, researchers studying these biological mediators should be aware of the strengths and weaknesses of the mechanistic concepts and methods used to measure them. They also should be aware of the untested assumptions about the clinical relevance of a number of proposed biological mechanisms (e.g., telomeres in nonreplicating cells, circulating cortisol). Another way to strengthen future study designs is to conduct experimental and quasiexperimental studies. Given that some environmental experiences cannot be assigned randomly for ethical reasons, intervention studies may be a useful way to evaluate changes that result when reducing environmental exposure to adversity. Some evidence suggests that interventions designed to address socioeconomic disadvantage may lead to improvements in physical health. In the 1990s, the United States Department of Housing and Urban Development undertook a large-scale randomized intervention study in five large U.S. cities to examine the health outcomes associated with moving out of poverty (Ludwig et al., 2011). Families were randomly assigned to one of three groups: One group received housing vouchers to move to a low-poverty area, another group received the vouchers without the low-poverty location restriction, and a third control group did not receive vouchers or counseling. Approximately 10 years later, Ludwig and colleagues measured a variety of metabolic outcomes, including glycosylated hemoglobin and BMI. Women who moved to low-poverty neighborhoods had decreased rates of extreme obesity and glycosylated hemoglobin relative to women in the control families. These findings suggest that reducing exposure to adverse neighborhood

experiences may have long-term health implications. This study was focused on health outcomes of adult women, but it is possible that similar intervention programs will have beneficial effects for children living in poverty. In fact, these types of intervention programs may have a stronger influence on children compared to adults if such an experience can create a lasting change in children's developmental trajectories.

Buffers and Protective Factors As described earlier in this chapter, epidemiological findings suggest that exposure to stressful adverse experiences is a significant risk factor for poor health outcomes. Nevertheless, not all individuals who experience adversity get sick (e.g., Chen et al., 2011; Miller, Lachman, et al., 2011). Research focused on buffers against the negative mental health outcomes associated with adversity has uncovered key variables that might help explain why some individuals thrive despite early experience with adversity. Researchers have identified intrapersonal characteristics (e.g., temperament, emotion regulation capacities), family characteristics (e.g., sensitive caregiving), and neighborhood factors (e.g., connections to adults in the community, access to support services) that serve as buffers and help protect children from some of the negative consequences associated with exposure to socioeconomic disadvantage and maltreatment. To date, most of this research has focused on protective factors for mental health outcomes, and only recently have researchers studied whether there might be buffers that mitigate the negative physical health outcomes associated with exposure to early adversity (Chen, Lee, Cavey, & Ho, 2013; Miller et al., 2011). For example, in a sample of healthy adolescents, low SES was associated with higher circulating markers of inflammation, but only for adolescents who did not have a supportive role model (Chen et al., 2013). Similarly, in a sample of adults who experienced low SES in childhood, high levels of maternal warmth buffered adults against a phenotype characterized by proinflammatory signaling (Chen et al., 2011). These findings are notable because they offer insight into the reasons why not all individuals exposed to early adversity go on to experience poor health in adulthood. Another related question concerns the timing of these buffers. Recent empirical evidence for the positive outcomes associated with protective factors has focused on buffers that are temporally connected to the experience of adversity (e.g., receiving supportive caregiving or having a role model while living in poverty). Research on the effects of these buffering factors also should examine the extent to which buffers at other important periods of development might play a role in mitigating risk. For example, adolescents who experience adversity in early childhood may benefit from supportive caregivers or role models in adolescence. It may be, however, that support systems need to be in place at the time of exposure to adversity to offset the negative experience. These questions will be particularly important to explore in order to shed light on what factors are most effective at ameliorating the negative health outcomes associated with exposure to adverse experiences.

Translational Implications

After identifying the protective factors that shield children from the negative health effects associated with early adversity, clinicians and researchers should use this information to develop testable interventions to examine whether the programs that foster behavioral changes also shape biological intermediaries in ways that have long-term implications for physical health. For example, programs designed to provide mentors and supportive role models to atrisk children may alter biological pathways that ultimately could improve long-term physical health outcomes. Similarly, parenting interventions that reduce children's exposure to harsh parental caregiving could shape biological processes in ways that minimize risk for accelerated aging and early mortality. For example, evidence from an intervention study suggests that children whose foster parents received a positive parenting intervention showed declines in basal cortisol levels relative to control children whose foster parents did not receive the intervention (Fisher, Gunnar, Chamberlain, & Reid, 2000). Although this study had a small sample size (just 10 children in each condition), the findings offer some indication that parenting interventions might affect predisease pathways in ways that are beneficial to health. Community programs that are designed to improve family socioeconomic conditions may provide opportunities for researchers to examine whether a modest change in the experience of early adversity shapes epigenetic modifications, HPA axis activity, and inflammation. These kinds of translational studies will require researchers and community advocates to work together to advance our understanding of how early adversity shapes physical health.

Linking Research on Developmental Psychopathology and Health Psychology Throughout this chapter, we have noted a number of instances in which exposure to early adversity has been associated with long-term mental health but not physical health outcomes (e.g., effects of exposure to interparental violence, natural disasters). This trend is not entirely surprising, given that programs of research often are focused on a narrow set of outcomes (e.g., mental health or physical health). Yet this tradition of focusing on a restricted range of outcomes ignores the fact that researchers interested in developmental psychopathology and health psychology often are motivated by a similar set of research goals. Going forward, we believe that a more integrated and interdisciplinary approach to research on mental and physical health is warranted. To the extent that researchers in these fields can share theories, methods, and work together collaboratively, scientific progress on all fronts will be much more rapid. Indeed, there are some exciting questions at the intersection of these fields, which we hope researchers will soon answer. For example, researchers across these disciplines frequently seek to understand the long-term developmental effects of early exposure to adversity and whether there are intrinsic or extrinsic buffers that can offset the negative outcomes associated with stress. These questions—and many others—will be of central interest in the years to come.

Conclusions We look forward to the next generation of studies on early adversity and adult physical health. In the coming years, more research on the biological intermediaries linking early adversity and

adult physical health will help uncover the processes by which stressful life events translate into chronic disease and accelerated aging. This research should take a developmental approach, using prospective studies and incorporating multiple levels of analysis, from molecular studies to large-scale epidemiological studies. The use of comprehensive models (e.g., models of equifinality and multifinality, developmental cascades, differential susceptibility) will help shed light on questions about (a) the relative timing of events in development (e.g., sensitive periods, transitional periods), (b) the complex interconnections among risk and protective factors, and (c) the biological consequences associated with early adversity. Findings from these studies will offer new ideas that can help inform public policy and clinical practice.

References Ackerman, S. H., Hofer, M. A., & Weiner, H. (1975). Age at maternal separation and gastric erosion susceptibility in the rat. Psychosomatic Medicine, 37, 180–184. Adam, E. K., & Kumari, M. (2009). Assessing salivary cortisol in large-scale, epidemiological research. Psychoneuroendocrinology, 34, 1423–1436. doi: 10.1016/j.psyneuen.2009.06.011 Ader, R. (1962). Social factors affecting emotionality and resistance to disease in animals: III. Early weaning and susceptibility to gastric ulcers in the rat: A control for nutritional factors. Journal of Comparative and Physiological Psychology, 55, 600–602. doi: 10.1037/h0041164 Ader, R., Tatum, R., & Beels, C. C. (1960). Social factors affecting emotionality and resistance to disease in animals: I. Age of separation from the mother and susceptibility to gastric ulcers in the rat. Journal of Comparative and Physiological Psychology, 53, 446–454. doi: 10.1037/h0044570 Adler, N. E., & Rehkopf, D. H. (2008). U.S. disparities in health: Descriptions, causes, and mechanisms. Annual Review of Public Health, 29, 235–252. doi: 10.1146/annurev.publhealth.29.020907.090852 Agardh, E. E., Ahlbom, A., Andersson, T., Efendic, S., Grill, V., Jallqvist, & Ostenson, C. G. (2007). Socio-economic position at three points in life in association with type 2 diabetes and impaired glucose tolerance in middle-aged Swedish men and women. International Journal of Epidemiology, 36, 84–92. doi: 10.1093/ije/dyl269 Anda, R. F., Felitti, V. J., Bremner, J. D., Walker, J. D., Whitfield, C., Perry, B. D.,… Giles, W. H. (2006). The enduring effects of abuse and related adverse experiences in childhood: A convergence of evidence from neurobiology and epidemiology. European Archives of Psychiatry and Clinical Neuroscience, 256, 174–186. doi: 10.1007/s00406–005–0624–4 Appleyard, K., Egeland, B., van Dulmen, M. H. M., & Sroufe, L. A. (2005). When more is not better: The role of cumulative risk in child behavior outcomes. Journal of Child Psychology

and Psychiatry, 46, 235–245. doi: 10.1111/j.1469–7610.2004.00351.x Asok, A., Bernard, K., Roth, T. L., Rosen, J. B., & Dozier, M. (2013). Parental responsiveness moderates the association between early-life stress and reduced telomere length. Development and Psychopathology, 25, 577–585. doi: 10.1017/S0954579413000011 Aubert, G., & Lansdorp, P. M. (2008). Telomeres and aging. Physiological Reviews, 88, 557– 579. doi: 10.1152/physrev.00026.2007 Avitsur, R., Hunzeker, J., & Sheridan, J. F. (2006). Role of early stress in the individual differences in host response to viral infection. Brain, Behavior, and Immunity, 20, 339–348. doi: 10.1016/j.bbi.2005.09.006 Aviv, A. (2008). The epidemiology of human telomeres: Faults and promises. Journal of Gerontology: Medical Sciences, 63A, 979–983. Aviv, A., Valdes, A. M., & Spector, T. D. (2006). Human telomere biology: Pitfalls of moving from the laboratory to epidemiology. International Journal of Epidemiology, 35, 1424–1429. doi: 10.1093/ije/dyl169 Bakermans-Kranenberg, M. J., & van IJzendoorn, M. H. (2011). Differential susceptibility to rearing environment depending on dopamine-related genes: New evidence and a metaanalysis. Development and Psychopathology, 23, 39–52. doi: 10.1017/S0954579410000635 Balagopal, P., de Ferranti, S. D., Cook, S., Daniels, S. R., Gidding, S. S., Hayman, L. L.,… Steinberger, J. (2011). Nontraditional risk factors and biomarkers for cardiovascular disease: Mechanistic, research, and clinical considerations for youth: A scientific statement from the American Heart Association. Circulation, 123, 2749–2769. doi: 10.1161/CIR.0b013e31821c7c64 Barker, D. J., & Clark, P. M. (1997). Fetal undernutrition and disease later in life. Reviews of Reproduction, 2, 105–112. doi: 10.1530/ror.0.0020105 Batten, S. V., Aslan, M., Maciejewski, P. K., & Mazure, C. M. (2004). Childhood maltreatment as a risk factor for adult cardiovascular disease and depression. Journal of Clinical Psychiatry, 65, 249–254. doi: 10.4088/JCP.v65n0217 Belsky, J. (1993). Etiology of child maltreatment: A developmental-ecological analysis. Psychological Bulletin, 114, 413–434. doi: 10.1037/0033–2909.114.3.413 Belsky, J. (1997). Theory testing, effect-size evaluation, and differential susceptibility to rearing influence: The case of mothering and attachment. Child Development, 68, 598–600. doi: 10.1111/j.1467–8624.1997.tb04221.x Belsky, J. (2005). Differential susceptibility to rearing influence. In B. J. Ellis & D. F. Bjorklund (Eds.), Origins of the social mind: Evolutionary psychology and child development (pp. 139–163). New York: Guilford Press.

Belsky, J., Bakermans-Kranenberg, M. J., & van IJzendoorn, M. H. (2007). For better and for worse: Differential susceptibility to environmental influences. Current Directions in Psychological Science, 16, 300–304. doi: 10.1111/j.1467–8721.2007.00525.x Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135, 885–908. doi: 10.1037/a0017376 Berenson, G. S., & Srinivasan, S. R. (2005). Cardiovascular risk factors in youth with implications for aging: The Bogalusa Heart Study. Neurobiology of Aging, 26, 303–307. doi: 10.1016/j.neurobiolaging.2004.05.009 Bertone-Johnson, E. R., Whitcomb, B. W., Missmer, S A., Karlson, E. W., & Rich-Edwards, J. W. (2012). Inflammation and early-life abuse in women. American Journal of Preventive Medicine, 43, 611–620. doi: 10.1016/j.amepre.2012.08.014 Bick, J., Naumova, O., Hunter, S., Barbot, B., Lee, M., Luthar, S. S.,… Grigorenko, E. L. (2012). Childhood adversity and DNA methylation of genes involved in the hypothalamuspituitary-adrenal axis and immune system: Whole-genome and candidate-gene associations. Development and Psychopathology, 24, 1417–1425. doi: 10.1017/S0954579412000806 Borghol, N., Suderman, M., McArdle, W., Racine, A., Hallett, M., Pembrey, M.,… Szyf, M. (2012). Associations with early-life socio-economic position in adult DNA methylation. International Journal of Epidemiology, 41, 62–74. doi: 10.1093/ije/dyr147 Bornstein, M. H., Hahn, C. S., & Haynes, O. M. (2010). Social competence, externalizing, and internalizing behavioral adjustment from early childhood through early adolescence: Developmental cascades. Development and Psychopathology, 22, 717–735. doi: 10.1017/S0954579410000416 Bowlby, J. (1969/1982). Attachment and loss: Vol. 1. Attachment. New York, NY: Basic Books. Bowlby, J. (1973). Attachment and loss: Vol 2. Separation. New York, NY: Basic Books. Bowlby, J. (1980). Attachment and loss: Vol. 3. Loss. New York, NY: Basic Books. Bowlby, J. (1988). A secure base: Parent-child attachment and healthy human development. New York, NY: Basic Books. Boyce, W. T., Chesney, M., Alkon, A., Tschann, J. M., Adams, S., Chesterman, B.,… Wara, D. (1995). Psychobiologic reactivity to stress and childhood respiratory illnesses: Results of two prospective studies. Psychosomatic Medicine, 57, 411–422. Boyce, W. T., & Ellis, B. J. (2005). Biological sensitivity to context: I. An evolutionarydevelopmental theory of the origins and functions of stress reactivity. Development & Psychopathology, 17, 271–301. doi: 10.1017/S0954579405050145 Brody, G. H., Yu, T., Chen, Y., Kogan, S. M., Evans, G. W., Windle, M.,… Philibert, R. A.

(2013). Supportive family environments, genes that confer sensitivity, and allostatic load among rural African American emerging adults: A prospective analysis. Journal of Family Psychology, 27, 22–29. doi: 10.1037/a0027829 Brown, D. W., Anda, R. F., Felitti, V. J., Edwards, V. J., Malarcher, A. M., Croft, J. B., & Giles, W. H. (2010). Adverse childhood experiences are associated with the risk of lung cancer: A prospective cohort study. BMC Public Health, 10, 20–32. doi: 10.1186/1471– 2458–10–20 Brunner, E. J., Marmot, M. G., Nanchahal, K., Shipley, M. J., Stansfeld, S. A., Juneja, M., & Alberti, K. G. M. M. (1997). Social inequality in coronary risk: Central obesity and the metabolic syndrome. Evidence from the Whitehall II study. Diabetologia, 40, 1341–1349. doi: 10.1007/s001250050830 Caldji, C., Tannenbaum, B., Sharma, S., Francis, D., Plotsky, P. M., & Meaney, M. J. (1998). Maternal care during infancy regulates the development of neural systems mediating the expression of fearfulness in the rat. Proceedings of the National Academy of Sciences, 95, 5335–5340. doi: 10.1073/pnas.95.9.5335 Campbell, D. T., & Fiske, D. W. (1959). Convergent and discriminant validation by the multitrait-multimethod matrix. Psychological Bulletin, 56, 81–105. doi: 10.1037/h0046016 Caspi, A., McClay, J., Moffitt, T. E., Mill, J., Martin, J., Craig, I. W.,… Poulton, R. (2002). Role of genotype in the cycle of violence in maltreated children. Science, 297, 851–854. doi: 10.1126/science.1072290 Chen, E., Lee, W. K., Cavey, L., & Ho, A. (2013). Role models and the psychological characteristics that buffer low-socioeconomic-status youth from cardiovascular disease. Child Development, 84, 1241–1252. doi: 10.1111/cdev.12037 Chen, E., Miller, G. E., Kobor, M. S., & Cole, S. W. (2011). Maternal warmth buffers the effects of low early-life socioeconomic status on pro-inflammatory signaling in adulthood. Molecular Psychiatry, 16, 729–737. doi: 10.1038/mp.2010.53 Chida, Y., & Steptoe, A. (2009). Cortisol awakening response and psychosocial factors: A systematic review and meta-analysis. Biological Psychology, 80, 265–278. doi: 10.1016/j.biopsycho.2008.10.004 Chida, Y., Sudo, N., Sonoda, J., Hiramoto, T., & Kubo, C. (2007). Early-life psychological stress exacerbates adult mouse asthma via the hypothalamus-pituitary-adrenal axis. American Journal of Respiratory and Critical Care, 175, 316–322. doi: 10.1164/rccm.200607–898OC Children's Defense Fund. (2012). The state of America's children. Washington, DC: Author. Cicchetti, D., & Dawson, G. (2002). Multiple levels of analysis. Development and Psychopathology, 14, 417–420. doi: 10.1017/S0954579402003012

Cicchetti, D., & Rogosch, F. A. (1996). Equifinality and multifinality in developmental psychopathology. Development and Psychopathology, 8, 597–600. doi: 10.1017/S0954579400007318 Cicchetti, D., & Toth, S. L. (2005). Child maltreatment. Annual Review of Clinical Psychology, 1, 409–438. doi: 10.1146/annurev.clinpsy.1.102803.144029 Claussen, B., Smith, G. D., & Thelle, D. (2003). Impact of childhood and adulthood socioeconomic position on cause specific mortality: The Oslo Mortality Study. Journal of Epidemiology & Community Health, 57, 40–45. doi: 10.1136/jech.57.1.40 Clow, A., Thorn, L., Evans, P., & Hucklebridge, F. (2004). The awakening cortisol response: Methodological issues and significance. Stress, 7, 29–37. doi: 10.1080/10253890410001667205 Cohen, S., Doyle, W. J., Turner, R. B., Alper, C. M., & Skoner, D. P. (2004). Childhood socioeconomic status and host resistance to infectious illness in adulthood. Psychosomatic Medicine, 66, 553–558. doi: 10.1097/01.psy.0000126200.05189.d3 Cohen, S., Janicki-Deverts, D., Chen, E., & Matthews, K. A. (2010). Childhood socioeconomic status and adult health. Annals of the New York Academy of Sciences, 1186, 37–55. doi: 10.1111/j.1749–6632.2009.05334.x Cohen, S., Janicki-Deverts, D., Turner, R. B., Casselbrant, M. L., Li-Korotky, H., Epel, E. S., & Doyle, W. J. (2013). Association between telomere length and experimentally induced upper respiratory viral infection in healthy adults. Journal of the American Medical Association, 309, 699–705. doi: 10.1001/jama.2013.613 Cole, S. W., Hawkley, L. C., Arevalo, J. M., Sung, C. S., Rose, R. M., & Cacioppo, J. T. (2007). Social regulation of gene expression in human leukocytes. Genome Biology, 8, R189. doi: 10.1186/gb-2007–8–9-r189 Cornier, M., Dabelea, D., Hernandez, T. L., Lindstrom, R. C., Steig, A. J., Stob, N. R.,… Eckel, R. H. (2008). The metabolic syndrome. Endocrine Reviews, 29, 777–822. doi: 10.1210/er.2008–0024 Cottrell, E. C., & Seckl, J. R. (2009). Prenatal stress, glucocorticoids and the programming of adult disease. Frontiers in Behavioral Neuroscience, 3, 1–19. doi: 10.3389/neuro.08.019.2009 Crouch, J. L., Hanson, R. F., Saunders, B. E., Kilpatrick, D. G., & Resnick, H. S. (2000). Income, race/ethnicity, and exposure to violence in youth: Results from the national survey of adolescents. Journal of Community Psychology, 28, 625–641. doi: 10.1002/1520– 6629(200011)28:63.0.CO;2-R Cutler, D. M., Miller, G., & Norton, D. M. (2007). Evidence on early-life income and late-life health from America's Dust Bowl era. Proceedings of the National Academy of Sciences, 104,

13244–13249. doi: 10.1073/pnas.0700035104 Damjanovic, A. K., Yang, Y., Glaser, R., Kiecolt-Glaser, J. K., Nguyen, H., Laskowski, B.,… Weng, N. P. (2007). Accelerated telomere erosion is associated with a declining immune function of caregivers of Alzheimer's disease patients. Journal of Immunology, 179, 4249– 4254. Danese, A., & McEwen, B. S. (2012). Adverse childhood experiences, allostasis, allostatic load, and age-related disease. Physiology & Behavior, 106, 29–39. doi: 10.1016/j.physbeh.2011.08.019 Danese, A., Moffitt, T. E., Harrington, H., Milne, B. J., Polanczyk, G., Pariante, C. M.,… Caspi, A. (2009). Adverse childhood experiences and adult risk factors for age-related disease: Depression, inflammation, and clustering of metabolic risk markers. Archives of Pediatric Adolescent Medicine, 163, 1135–1143. doi: 10.1001/archpediatrics.2009.214 Danese, A., Pariante, C. M., Caspi, A., Taylor, A., & Poulton, R. (2007). Childhood maltreatment predicts adult inflammation in a life-course study. Proceedings of the National Academy of Sciences, 104, 1319–1324. doi: 10.1073/pnas.0610362104 Danesh, J., Collins, R., Appleby, P., & Peto, R. (1998). Association of fibrinogen, C-reactive protein, albumin, or leukocyte count with coronary heart disease: Meta-analyses of prospective studies. Journal of the American Medical Association, 279, 1477–1482. doi: 10.1001/jama.279.18.1477 Datar, A., Liu, J., Linnemayr, S., & Stecher, C. (2013). The impact of natural disasters on child health and investments in rural India. Social Science & Medicine, 76, 83–91. doi: 10.1016/j.socscimed.2012.10.008 De Lange, T. (2009). How telomeres solve the end-protection problem. Science, 326, 948– 952. doi: 10.1126/science.1170633 de Rooij, S. R., Wouters, H., Yonker, J. E., Painter, R. C., & Roseboom, T. J. (2010). Prenatal undernutrition and cognitive function in late adulthood. Proceedings of the National Academy of Sciences, 107, 16881–16886. doi: 10.1073/pnas.1009459107 DeVol, R., & Bedroussian, A. (2007). An unhealthy America: The economic burden of chronic disease: Charting a new course to save lives and increase productivity and economic growth. Santa Monica, CA: Milken Institute. Dickerson, S. S., & Kemeny, M. E. (2004). Acute stressors and cortisol responses: A theoretical integration and synthesis of laboratory research. Psychological Bulletin, 130, 355– 391. doi: 10.1037/0033–2909.130.3.355 Dill, D. L., Chu, J. A., Grob, M. C., & Eisen, S. V. (1991). The reliability of abuse history reports: A comparison of two inquiry formats. Comprehensive Psychiatry, 32, 166–169. doi: 10.1016/0010–440X(91)90009–2

Dong, M., Giles, W. H., Felitti, V. J., Dube, S. R., Williams, J. E., Chapman, D. P., & Anda, R. F. (2004). Insights into causal pathways for ischemic heart disease: Adverse childhood experiences study. Circulation, 110, 1761–1766. doi: 10.1161/01.CIR.0000143074.54995.7F Draper, B., Pfaff, J. J., Pirkis, J., Snowdon, J., Lautenschlager, N. T., Wilson, I.,… Almeida, O. P. (2008). Long-term effects of childhood abuse on the quality of life and health of older people: Results from the depression and early prevention of suicide in general practice project. Journal of the American Geriatrics Society, 56, 262–271. doi: 10.1111/j.1532– 5415.2007.01537.x Drury, S. S., Theall, K., Gleason, M. M., Smyke, A. T., DeVivo, I., Wong, J. Y. Y.,… Nelson, C. A. (2012). Telomere length and early severe social deprivation: Linking early adversity and cellular aging. Molecular Psychiatry, 17, 719–727. doi: 10.1038/mp.2011.53

Duncan, G. J., Kalil, A., & Ziol-Guest, K. M. (2008). The economic costs of early childhood poverty. Retrieved from http://www.pewtrusts.org/uploadedFiles/wwwpewtrustsorg/Reports/Partnership_for_Americas_Econ Ellis, B. J., Boyce, W. T., Belsky, J., Bakermans-Kranenberg, M. J., & van IJzendoorn, M. H. (2011). Differential susceptibility to the environment: An evolutionary-neurodevelopmental theory. Development and Psychopathology, 23, 7–28. doi: 10.1017/S0954579410000611 Epel, E. S. (2009). Telomeres in a life-span perspective: A new “psychobiomarker”? Current Directions in Psychological Science, 18, 6–10. doi: 10.1111/j.1467–8721.2009.01596.x Essex, M. J., Boyce, T. W., Hertzman, C., Lam, L. L., Armstrong, J. M., Neumann, S. M., & Kobor, M. S. (2013). Epigenetic vestiges of early developmental adversity: Childhood stress exposure and DNA methylation in adolescence. Child Development, 84, 58–75. doi: 10.1111/j.1467–8624.2011.01641.x Evans, G. W. (2004). The environment of childhood poverty. American Psychologist, 59, 77– 92. doi: 10.1037/0003–066X.59.2.77 Evans, G. W., & Kim, P. (2012). Childhood poverty and young adults' allostatic load: The mediating role of childhood cumulative risk exposure. Psychological Science, 23, 979–983. doi: 10.1177/0956797612441218 Fang, X., Brown, D. S., Florence, C. S., & Mercy, J. A. (2012). The economic burden of child maltreatment in the United States and implications for prevention. Child Abuse & Neglect, 36, 156–165. doi: 10.1016/j.chiabu.2011.10.006 Feldman, P. J., Cohen, S., Doyle, W. J., Skoner, D. P., & Gwaltney, J. M. (1999). The impact of personality on the reporting of unfounded symptoms and illnesses. Journal of Personality and Social Psychology, 77, 370–378. doi: 10.1037/0022–3514.77.2.370 Felitti, V. J., Anda, R. F., Nordenberg, D., Williams, D. F., Spitz, A. M., Edwards, V.,… Marks, J. S. (1998). Relationship of childhood abuse and household dysfunction to many of the leading

causes of death in adults: The Adverse Childhood Experiences (ACE) Study. American Journal of Preventive Medicine, 14, 245–258. doi: 10.1016/S0749- 3797(98)00017–8 Field, T., Diego, M., Dieter, J., Hernandez-Reif, M., Schanberg, S., Kuhn, C.,… Bendell, D. (2004). Prenatal depression effects on the fetus and the newborn. Infant Behavior and Development, 27, 216–229. doi: 10.1016/j.infbeh.2003.09.010 Fisher, P. A., Gunnar, M. R., Chamberlain, P., & Reid, J. B. (2000). Preventative intervention for maltreated preschool children: Impact on children's behavior, neuroendocrine activity, and foster parent functioning. Journal of the American Academy of Child and Adolescent Psychiatry, 39, 1356–1364. doi: 10.1097/00004583–200011000–00009 Fuller-Thomson, E., Bejan, R., Hunter, J. T., Grundland, T., & Brennenstuhl, S. (2012). The link between childhood sexual abuse and myocardial infarction in a population-based study. Child Abuse & Neglect, 36, 656–665. doi: 10.1016/j.chiabu.2012.06.001 Fuller-Thomson, E., & Brennenstuhl, S. (2009). Making a link between childhood abuse and cancer: Results from a regional representative survey. Cancer, 115, 3341–3350. doi: 10.1002/cncr.24372 Fuller-Thomson, E., Brennenstuhl, S., & Frank, J. (2010). The association between childhood physical abuse and heart disease in adulthood: Findings from a representative community sample. Child Abuse & Neglect, 34, 689–698. doi: 10.1016/j.chiabu.2010.02.005 Gadalla, S. M., Cawthon, R., Giri, N., Alter, B. P., & Savage, S. A. (2010). Telomere length in blood, buccal cells, and fibroblasts from patients with inherited bone marrow failure syndromes. Aging, 2, 867–874. Galobardes, B., Lynch, J. W., & Smith, G. D. (2004). Childhood socioeconomic circumstances and cause-specific mortality in adulthood: Systematic review and interpretation. Epidemiologic Reviews, 26, 7–21. doi: 10.1093/epirev/mxh008 Galobardes, B., Lynch, J. W., & Smith, G. D. (2008). Is the association between childhood socioeconomic circumstances and cause-specific mortality established? Update of a systematic review. Journal of Epidemiology & Community Health, 62, 387–390. doi: 10.1136/jech.2007.065508 Galobardes, B., Smith, G. D., & Lynch, J. W. (2006). Systematic review of the influence of childhood socioeconomic circumstances on risk for cardiovascular disease in adulthood. Annals of Epidemiology, 16, 91–104. doi: 10.1016/j.annepidem.2005.06.053 Geronimus, A. T., Hicken, M. T., Pearson, J. A., Seashols, S. J., Brown, K. L., & Cruz, T. D. (2010). Do U.S. Black women experience stress-related accelerated biological aging? A novel theory and first population-based test of Black-White differences in telomere length. Human Nature, 21, 19–38. doi: 10.1007/s12110–010–9078–0 Gerritson, L., Geerlings, M. I., Beekman, A. T. F., Deeg, D. J. H., Penninx, B. W. J. H., &

Comijs, H. C. (2010). Early and late life events and salivary cortisol in older persons. Psychological Medicine, 40, 1569–1578. doi: 10.1017/S0033291709991863 Gluckman, P. D., & Hanson, M. A. (2006). The conceptual basis for developmental origins of health and disease. In P. Gluckman & M. Hanson (Eds.), Developmental origins of health and disease (pp. 33–50). New York, NY: Cambridge University Press. Goodwin, R. D., & Stein, M. B. (2004). Association between childhood trauma and physical disorders among adults in the United States. Psychological Medicine, 34, 509–520. doi: 10.1017/S003329170300134X Graber, J. A., & Brooks-Gunn, J. (1996). Transitions and turning points: Navigating the passage from childhood through adolescence. Developmental Psychology, 32, 768–776. doi: 10.1037/0012–1649.32.4.768 Gruenewald, T. L., Karlamangla, A. S., Hu, P., Stein-Merkin, S., Crandall, C., Koretz, B., & Seeman, T. E. (2012). History of socioeconomic disadvantage and allostatic load in later life. Social Science & Medicine, 74, 75–83. doi: 10.1016/j.socscimed.2011.09.037 Grundy, S. M., Cleeman, J. I., Daniels, S. R., Donato, K. A., Eckel, R. H., Franklin, B. A.,… Costa, F. (2005). Diagnosis and management of the metabolic syndrome: An American Heart Association/National Heart, Lung, and Blood Institute scientific statement. Circulation, 112, 2735–2752. doi: 10.1161/CIRCULATIONAHA.105.169404 Gunnar, M. R., & Vazquez, D. M. (2001). Low cortisol and a flattening of expected daytime rhythm: Potential indices of risk in human development. Development and Psychopathology, 13, 515–538. doi: 10.1017/S0954579401003066 Gustafsson, P. E., Janlert, U., Theorell, T., Westerlund, H., & Hammarstrom, A. (2012). Social and material adversity from adolescence to adulthood and allostatic load in middle-aged women and men: Results from the Northern Swedish Cohort. Annals of Behavioral Medicine, 43, 117–128. doi: 10.1007/s12160–011–9309–6 Hammen, C., Henry, R., & Daley, S. E. (2000). Depression and sensitization to stressors among young women as a function of childhood adversity. Journal of Consulting and Clinical Psychology, 68, 782–787. doi: 10.1037/0022–006X.68.5.782 Heijmans, B. T., Tobi, E. W., Stein, A. D., Putter, H., Blauw, G. J., Susser, E. S.,… Lumey, L. H. (2008). Persistent epigenetic differences associated with prenatal exposure to famine in humans. Proceedings of the National Academy of Sciences, 105, 17046–17049. doi: 10.1073/pnas.0806560105 Heim, C., Ehlert, U., & Hellhammer, D. H. (2000). The potential role of hypocortisolism in the pathophysiology of stress-related bodily disorders. Psychoneuroendocrinology, 25, 1–35. doi: 10.1016/S0306–4530(99)00035–9 Herrenkohl, T. I., Kosterman, R., Mason, W. A., Hawkins, J. D., McCarty, C. A., & McCauley,

E. (2010). Effects of childhood conduct problems and family adversity on health, health behaviors, and service use in early adulthood: Tests of developmental pathways involving adolescent risk taking and depression. Development and Psychopathology, 22, 655–665. doi: 10.1017/S0954579410000349 Hertzman, C. (1999). The biological embedding of early experience and its effects on health in adulthood. Annals of the New York Academy of Sciences, 896, 85–95. doi: 10.1111/j.1749– 6632.1999.tb08107.x Hofer, M. A. (2002). The riddle of development. In D. J. Lewkowicz & R. Lickliter (Eds.), Conceptions of Development (pp. 5–30). Philadelphia, PA: Psychology Press. Holmbeck, G. N. (1994). Adolescence. In V. S. Ramachandran (Ed.), Encyclopedia of human behavior (Vol. 1, pp. 17–28). Orlando, FL: Academic Press. Hucklebridge, F. H., Mellins, J., Evans, P., & Clow, A. (2002). The awakening cortisol response: No evidence for an influence on body posture. Life Sciences, 71, 639–646. doi: 10.1016/S0024–3205(02)01726–5 Ingram, R. E., & Luxton, D. D. (2005). Vulnerability-stress models. In B. L. Hankin & J. R. Z. Abela (Eds.), Development of psychopathology: A vulnerability-stress perspective (pp. 32– 46). Thousand Oaks, CA: Sage. Irish, L., Kobayashi, I., & Delahanty, D. L. (2010). Long-term physical health consequences of childhood sexual abuse: A meta-analytic review. Journal of Pediatric Psychology, 35, 450– 461. doi: 10.1093/jpepsy/jsp118 Jacobs, J. R., & Bovasso, G. B. (2000). Early and chronic stress and their relation to breast cancer. Psychological Medicine, 30, 669–678. Jaenisch, R., & Bird, A. (2003). Epigenetic regulation of gene expression: How the genome integrates intrinsic and environmental signals. Nature Genetics, 33, 245–254. doi: 10.1038/ng1089 Jirtle, R. L., & Skinner, M. K. (2007). Environmental epigenomics and disease susceptibility. Nature Reviews: Genetics, 8, 253–262. doi: 10.1038/nrg2045 Johnson, R. C., & Schoeni, R. F. (2011). Early-life origins of adult disease: National longitudinal population-based study of the United States. American Journal of Public Health, 101, 2317–2324. doi: 10.2105/AJPH.2011.300252 Kananen, L., Surakka, I., Pirkola, S., Suvisaari, J., Lönnqvist, J., Peltonen, L.,… Hovatta, I. (2010). Childhood adversities are associated with shorter telomere length at adult age both in individuals with an anxiety disorder and controls. Plos One, 5, e10826. doi: 10.1371/journal.pone.0010826 Kario, K., McEwen, B. S., & Pickering, T. G. (2003). Disasters and the heart: A review of the

effects of earthquake-induced stress on cardiovascular disease. Hypertension Research, 26, 355–367. Karlamangla, A. S., Singer, B. H., & Seeman, T. E. (2006). Reduction in allostatic load in older adults is associated with lower all-cause mortality risk: MacArthur studies of successful aging. Psychosomatic Medicine, 68, 500–507. doi: 10.1097/01.psy.0000221270.93985.82 Keinan-Boker, L., Vin-Raviv, N., Liphshitz, I., Linn, S., & Barchana, M. (2009). Cancer incidence in Israeli Jewish survivors of World War II. Journal of the National Cancer Institute, 101, 1489–1500. doi: 10.1093/jnci/djp327 Kessler, R. C., Davis, C. G., & Kendler, K. S. (1997). Childhood adversity and adult psychiatry disorder in the U.S. National Comorbidity Survey. Psychological Medicine, 27, 1101–1119. doi: 10.1017/S0033291797005588 Kiecolt-Glaser, J. K., Gouin, J. P., Weng, N. P., Malarkey, W. B., Beversdorf, D. Q., & Glaser, R. (2011). Childhood adversity heightens the impact of later-life caregiving stress on telomere length and inflammation. Psychosomatic Medicine, 73, 16–22. doi: 10.1097/PSY.0b013e31820573b6 Kimura, M., Hjelmborg, J. V., Gardner, J. P., Bathum, L., Brimacombe, M., Lu, X.,… Christensen, K. (2008). Telomere length and mortality: A study of leukocytes in elderly Danish twins. American Journal of Epidemiology, 167, 799–806. doi: 10.1093/aje/kwm380 Kirschbaum, C., & Hellhammer, D. H. (1989). Salivary cortisol in psychobiological research: An overview. Neuropsychobiology, 22, 150–169. doi: 10.1159/000118611 Kittleson, M. M., Meoni, L. A., Wang, N. Y., Chu, A. Y., Ford, D. E., & Klag, M. J. (2006). Association of childhood socioeconomic status with subsequent coronary heart disease in physicians. Archives of Internal Medicine, 166, 2356–2361. doi: 10.1001/archinte.166.21.2356 Kivelä, S. L., Luukinen, H., Koski, K., Viramo, P., & Kimmo, P. (1998). Early loss of mother or father predicts depression in old age. International Journal of Geriatric Psychiatry, 13, 527–530. doi: 10.1002/(SICI)1099–1166(199808)13:83.0.CO;2–8 Knudsen, E. I. (2004). Sensitive periods in the development of the brain and behavior. Journal of Cognitive Neuroscience, 16, 1412–1425. doi: 10.1162/0898929042304796 Korpimäki, S. K., Sumanen, M. P., Silanmäki,L. H., & Mattila, K. J. (2010). Cancer in working-age is not associated with childhood adversities. Acta Oncologica, 49, 436–440. doi: 10.3109/02841860903521103 Krause, N. (1998). Early parental loss, recent life events, and changes in health among older adults. Journal of Aging and Health, 10, 395–421. doi: 10.1177/089826439801000401 Kruschinski, C., Skripuletz, T., Bedoui, S., Raber, K., Straub, R. H., Hoffmann, T.,… von

Horsten, S. (2008). Postnatal life events affect the severity of asthmatic airway inflammation in the adult rat. Journal of Immunology, 180, 3919–3925. Kumari, M., Head, J., Bartley, M., Stansfeld, S., & Kivimaki, M. (2013). Maternal separation in childhood and diurnal cortisol patterns in mid-life: Findings from the Whitehall II study. Psychological Medicine, 43, 633–643. doi: 10.1017/S0033291712001353 Kumari, M., Head, J., & Marmot, M. (2004). Prospective study of social and other risk factors for incidence of type 2 diabetes in the Whitehall II study. Archives of Internal Medicine, 164, 1873–1880. doi: 10.1001/archinte.164.17.1873 Lam, L. L., Emberly, E., Fraser, H. B., Neumann, S. M., Chen, E., Miller, G. E., & Kobor, M. S. (2012). Factors underlying variable DNA methylation in a human community cohort. Proceedings of the National Academy of Sciences, 109, 17253–17260. doi: 10.1073/pnas.1121249109 Lawlor, D. A., Smith, G. D., Rumley, A., Lowe, G. D. O., & Ebrahim, S. (2005). Associations of fibrinogen and C-reactive protein with prevalent and incident coronary heart disease are attenuated by adjustment for confounding factors: British Women's Heart and Health Study. Thrombosis and Haemostasis, 93, 955–963. doi: 10.1160/TH04–12–0805 Lawlor, D. A., Sterne, J. A. C., Tynelius, P., Smith, G. D., Rasmussen, F. (2006). Association of childhood socioeconomic position with cause-specific mortality in a prospective record linkage study of 1,839,384 individuals. American Journal of Epidemiology, 164, 907–915. doi: 10.1093/aje/kwj319 Lerner, R. M., & Castellino, D. R. (2002). Contemporary developmental theory and adolescence: Developmental systems and applied developmental science. Journal of Adolescent Health, 31, 122–135. doi: 10.1016/S1054–139X(02)00495–0 Li, L., Power, C., Kelly, S., Kirschbaum, C., & Hertzman, C. (2007). Life-time socio-economic position and cortisol patterns in mid-life. Psychoneuroendocrinology, 32, 824–833. doi: 10.1016/j.psyneuen.2007.05.014 Libby, P., Ridker, P. M., & Hansson, G. K. (2009). Inflammation in atherosclerosis: From pathophysiology to practice. Journal of the American College of Cardiology, 54, 2129–2138. doi: 10.1016/j.jacc.2009.09.009 Lindblad, U., Langer, R. D., Wingard, D. L., Thomas, R. G., & Barrett-Connor, E. L. (2001). Metabolic syndrome and ischemic heart disease in elderly men and women. American Journal of Epidemiology, 153, 481–489. doi: 10.1093/aje/153.5.481 Liu, D., Diorio, J., Tannenbaum, B., Caldji, C., Francis, D., Freedman, A, Sharma, S.,… Meaney, M. (1997). Maternal care, hippocampal glucocorticoid receptor gene expression and hypothalamic-pituitary-adrenal responses to stress. Science, 277, 1659–1662. doi: 10.1126/science.277.5332.1659

Lucas, A., Fewtrell, M. S., & Cole, T. J. (1999). Fetal origins of adult disease – the hypothesis revisited. British Medical Journal, 319, 245–249. doi: 10.1136/bmj.319.7204.245 Ludwig, J., Sanbonmatsu, L., Gennetian, L., Adam, E., Duncan, G. J., Katz, L. F.,…& McDade, T. W. (2011). Neighborhoods, obesity, and diabetes – A randomized social experiment. New England Journal of Medicine, 365, 1509–1519. doi: 10.1056/NEJMsa1103216 Luecken, L. J., & Roubinov, D. S. (2012). Pathways to lifespan health following childhood parental death. Social and Personality Psychology Compass, 6, 243–257. doi: 10.1111/j.1751–9004.2011.00422.x Lule, E., Rosen, J. E., Singh, S., Knowles, J. C., & Behrman, J. R. (2006). Adolescent health programs. In D. T. Jamison, J. G. Breman, A. R. Measham, et al. (Eds.), Disease control priorities in developing countries (2nd ed., pp. 217–234). Washington, DC: World Bank. Lynch, J., & Smith, G. D. (2005). A life course approach to chronic disease epidemiology. Annual Review of Public Health, 26, 1–35. doi: 10.1146/annurev.publhealth.26.021304.144505 Lyons-Ruth, K., & Jacobvitz, D. (2008). Attachment disorganization: Genetic factors, parenting contexts, and developmental transformation from infancy to adulthood. In J. Cassidy & P. R. Shaver (Eds.), Handbook of attachment: Theory, research, and clinical applications (2nd ed., pp. 666–697). New York, NY: Guilford Press. Maier, E. H., & Lachman, M. E. (2000). Consequences of early parental loss and separation for health and well-being in midlife. International Journal of Behavioral Development, 24, 183–189. doi: 10.1080/016502500383304 Masten, A. S., Burt, K., & Coatsworth, J. D. (2006). Competence and psychopathology in development. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology (Vol. 3, 2nd ed., pp. 696–738). Hoboken, NJ: Wiley. Masten, A. S., & Cicchetti, D. (2010). Developmental cascades. Development and Psychopathology, 22, 491–495. doi: 10.1017/S0954579410000222 Mather, K. A., Jorm, A. F., Parslow, R. A., & Christensen, H. (2011). Is telomere length a biomarker of aging? A review. Journal of Gerontology, 66A, 202–213. doi: 10.1093/gerona/glq180 Matthews, K. A., & Gallo, L. C. (2011). Psychological perspectives on pathways linking socioeconomic status and physical health. Annual Review of Psychology, 62, 501–530. doi: 10.1146/annurev.psych.031809.130711 McCaffery, J. M., Marsland, A. L., Strohacker, K., Muldoon, M. F., & Manuck, S. B. (2012). Factor structure underlying components of allostatic load. Plos One, 7, e47246. doi: 10.1371/journal.pone.0047246

McEwen, B. S. (1998). Protective and damaging effects of stress mediators. New England Journal of Medicine, 338, 171–179. doi: 10.1056/NEJM199801153380307 McEwen, B. S., & Seeman, T. (1999). Protective and damaging effects of mediators of stress: Elaborating and testing the concepts of allostasis and allostatic load. Annals of the New York Academy of Sciences, 896, 30–47. doi: 10.1111/j.1749–6632.1999.tb08103.x McEwen, B. S., & Stellar, E. (1993). Stress and the individual: Mechanisms leading to disease. Archives of Internal Medicine, 153, 2093–2101. doi: 10.1001/archinte.1993.00410180039004 McGowan, P. O., Sasaki, A., D'Alessio, A. C., Dymov, S., Labonte, B., Szyf, M.,… Meaney, M. J. (2009). Epigenetic regulation of the glucocorticoid receptor in human brain associates with child abuse. Nature Neuroscience, 12, 342–348. doi: 10.1038/nn.2270 Meaney, M. J. (2001). Maternal care, gene expression, and the transmission of individual differences in stress reactivity across generations. Annual Reviews Neuroscience, 24, 1161– 1192. doi: 10.1146/annurev.neuro.24.1.1161 Meaney, M. J., & Aitken, D. H. (1985). The effects of early postnatal handling on hippocampal glucocorticoid receptor concentrations: Temporal parameters. Brain Research, 354, 301–304. doi: 10.1016/0165–3806(85)90183-X Melchior, M., Moffitt, T. E., Milne, B. J., Poulton, R., & Caspi, A. (2007). Why do children from socioeconomically disadvantaged families suffer from poor health when they reach adulthood? A life-course study. American Journal of Epidemiology, 166, 966–974. doi: 10.1093/aje/kwm155 Midei, A. J., Matthews, K. A., Chang, Y., & Bromberger, J. T. (2013). Childhood physical abuse is associated with incident metabolic syndrome in mid-life women. Health Psychology, 32, 121–127. doi: 10.1037/a0027891 Miller, G. E., Chen, E., Fok, A. K., Walker, H., Lim, A., Nicholls, E. F.,… Kobor, M. S. (2009). Low early-life social class leaves a biological residue manifested by decreased glucocorticoid and increased proinflammatory signaling. Proceedings of the National Academy of Sciences, 106, 14716–14721. doi: 10.1073/pnas.0902971106 Miller, G. E., Chen, E., & Parker, K. J. (2011). Psychological stress in childhood and susceptibility to the chronic diseases of aging: Moving toward a model of behavioral and biological mechanisms. Psychological Bulletin, 137, 959–997. doi: 10.1037/a0024768 Miller, G. E., Chen, E., & Zhou, E. S. (2007). If it goes up, must it come down? Chronic stress and the hypothalamic-pituitary-adrenocortical axis in humans. Psychological Bulletin, 133, 25–45. doi: 10.1037/0033–2909.133.1.25 Miller, G. E., Lachman, M. E., Chen, E., Gruenewald, T. L., Karlamangla, A. S., & Seeman, T. E. (2011). Pathways to resilience: Maternal nurturance as a buffer against the effects of

childhood poverty on metabolic syndrome at midlife. Psychological Science, 22, 1591–1599. doi: 10.1177/0956797611419170 Moffitt, T. E., Caspi, A., & Rutter, M. (2006). Measured gene-environment interactions in psychopathology: Concepts, research strategies, and implications for research, intervention and public understanding of genetics. Perspectives on Psychological Science, 1, 5–27. doi: 10.1111/j.1745–6916.2006.00002.x Morton, P. M., Schafer, M. H., & Ferraro, K. F. (2012). Does misfortune increase cancer risk in adulthood? Journal of Aging and Health, 24, 948–984. doi: 10.1177/0898264312449184 Murphy, S. L., Xu, J., & Kochanek, K. D. (2012). Deaths: Preliminary data for 2010. National Vital Statistics Reports, Vol. 60, no. 4. Hyattsville, MD: National Center for Health Statistics. Nandi, A., Glymour, M. M., Kawachi, I., & VanderWeele, T. J. (2012). Using marginal structural models to estimate the direct effect of adverse childhood social conditions on onset of heart disease, diabetes, and stroke. Epidemiology, 23, 223–232. doi: 10.1097/EDE.0b013e31824570bd Nicolson, N. A. (2004). Childhood parental loss and cortisol levels in adult men. Psychoneuroendocrinology, 29, 1012–1018. doi: 10.1016/j.psyneuen.2003.09.005 Oberlander, T., Weinberg, J., Papsdorf, M., Grunau, R., Misri, S., & Devlin, A. (2008). Prenatal exposure to maternal depression, neonatal methylation of human glucocorticoid receptor gene (NR3C1) and infant cortisol stress responses. Epigenetics, 3, 97–106. doi: 10.4161/epi.3.2.6034 Ohgane, J., Yagi, S., & Shiota, K. (2008). Epigenetics: The DNA methylation profile of tissuedependent and differentially methylated regions in cells. Placenta, 29 (Suppl. A), S29–35. doi: 10.1016/j.placenta.2007.09.011 O'Sullivan, L., Combes, A. N., & Moritz, K. M. (2012). Epigenetics and developmental programming of adult onset diseases. Pediatric Nephrology, 27, 2175–2182. doi: 10.1007/s00467–012–2108-x Painter, R. C., de Rooij, S. R., Bossuyt, P. M., Simmers, T. A., Osmond, C., Barker, D. J.,… Roseboom, T. J. (2006). Early onset of coronary artery disease after prenatal exposure to the Dutch famine. American Journal of Clinical Nutrition, 84, 322–327. Painter, R. C., Roseboom, T. J., & Bleker, O. P. (2005). Prenatal exposure to the Dutch famine and disease in later life: An overview. Reproductive Toxicology, 20, 345–352. doi: 10.1016/j.reprotox.2005.04.005 Parker, L., Lamont, D. W., Unwin, N., Pearce, M. S., Bennett, S. M. A., Dickinson, H. O.,… Craft, A. W. (2003). A lifecourse study of risk for hyperinsulinaemia, dyslipidaemia and obesity (the central metabolic syndrome) at age 49–51 years. Diabetic Medicine, 20, 406– 415. doi: 10.1046/j.1464–5491.2003.00949.x

Pesonen, A., Raikkonen, K., Feldt, K., Heinonen, K., Osmond, C., Phillips, D. I. W.,… Kajantie, E. (2010). Childhood separation experience predicts HPA axis hormonal responses in late adulthood: A natural experiment of World War II. Psychoneuroendocrinology, 35, 758– 767. doi: 10.1016/j.psyneuen.2009.10.017 Pesonen, A., Raikkonen, K., Heinonen, K., Kajantie, E., Forsen, T., & Eriksson, J. G. (2007). Depressive symptoms in adults separated from their parents as children: Natural experiment during World War II. American Journal of Epidemiology, 166, 1126–1133. doi: 10.1093/aje/kwm254 Pollitt, R. A., Kaufman, J. S., Rose, K. M., Diez-Roux, A. V., Zeng, D., & Hess, G. (2007). Early-life and adult socioeconomic status and inflammatory risk markers in adulthood. European Journal of Epidemiology, 22, 55–66. doi: 10.1007/s10654–006–9082–1 Poulton, R., Caspi, A., Milne, B. J., Thomson, W. M., Taylor, A., Sears, M. R., & Moffitt, T. E. (2002). Association between children's experience of socioeconomic disadvantage and adult health: A life-course study. Lancet, 360, 1640–1645. doi: 10.1016/S0140–6736(02)11602–3 Power, C., Hyppönen, E., & Smith, G. D. (2005). Socioeconomic position in childhood and early adult life and risk of mortality: A prospective study of the mothers of the 1958 British birth cohort. American Journal of Public Health, 95, 1396–1402. doi: 10.2105/AJPH.2004.047340 Price, L. H., Kao, H., Burgers, D. E., Carpenter, L. L., & Tyrka, A. R. (2013). Telomeres and early-life stress: An overview. Biological Psychiatry, 73, 15–23. doi: 10.1016/j.biopsych.2012.06.025 Pudrovska, T., Anishkin, A., & Shen, Y. (2012). Early-life socioeconomic status and the prevalence of breast cancer in later life. Research on Aging, 34, 302–320. doi: 10.1177/0164027511415632 Rabin, B. S. (2005). Introduction to immunology and immune-endocrine interactions. In K. Vedhara & M. R. Irwin (Eds.), Human psychoneuroimmunology (pp. 1–24). New York, NY: Oxford University Press. Raison, C. L., & Miller, A. H. (2003). When not enough is too much: The role of insufficient glucocorticoid signaling in the pathophysiology of stress-related disorders. American Journal of Psychiatry, 160, 1554–1565. doi: 10.1176/appi.ajp.160.9.1554 Rakyan, V. K., Down, T. A., Balding, D. J., & Beck, S. (2011). Epigenome-wide association studies for common human diseases. Nature Reviews Genetics, 12, 529–541. doi: 10.1038/nrg3000 Repetti, R. L., Robles, T. F., & Reynolds, B. (2011). Allostatic processes in the family. Development and Psychopathology, 23, 921–938. doi: 10.1017/S095457941100040X Repetti, R. L., Taylor, S. E., & Seeman, T. E. (2002). Risky families: Family social

environments and the mental and physical health of offspring. Psychological Bulletin, 128, 330–366. doi: 10.1037/0033–2909.128.2.330 Rich-Edwards, J. W., Colditz, F. A., Stampfer, M. J., Willett, W. C., Gillman, M. W., Hennekens, C. H.,… Manson, J. E. (1999). Birthweight and the risk for type 2 diabetes mellitus in adult women. Annals of Internal Medicine, 130, 278–284. doi: 10.7326/0003– 4819–130–4_Part_1–199902160–00005 Rich-Edwards, J. W., Mason, S., Rexrode, K., Spiegelman, D., Hibert, E., Kawachi, I.,… Wright, R. J. (2012). Physical and sexual abuse in childhood as predictors of early-onset cardiovascular events in women. Circulation, 126, 920–927. doi: 10.1161/CIRCULATIONAHA.111.076877 Rich-Edwards, J. W., Spiegelman, D., Hibert, E. N. L., Jun, H. J., Todd, T. J., Kawachi, I., & Wright, R. J. (2010). Abuse in childhood and adolescence as a predictor of type 2 diabetes in adult women. American Journal of Preventive Medicine, 39, 529–536. doi: 10.1016/j.amepre.2010.09.007 Ridker, P. M. (2007). Inflammatory biomarkers and risks of myocardial infarction, stroke, diabetes, and total mortality: Implications for longevity. Nutrition Reviews, 65, S253–259. doi: 10.1111/j.1753–4887.2007.tb00372.x Romans, S., Belaise, C., Martin, J., Morris, E., & Raffi, A. (2002). Childhood abuse and later medical disorders in women: An epidemiological study. Psychotherapy and Psychosomatics, 71, 141–150. doi: 10.1159/000056281 Ross, C. E., & Mirowsky, J. (2001). Neighborhood disadvantage, disorder, and health. Journal of Health and Social Behavior, 42, 258–276. Ross, K. M., Murphy, M. L. M., Adam, E. K., Chen, E., & Miller, G. E. (2014). How stable are diurnal cortisol activity indices in healthy individuals? Evidence from three multi-wave studies. Psychoneuroendocrinology, 39, 184–193. doi: 10.1016/j.psyneuen.2013.09.016 Roth, T. L., Lubin, F. D., Funk, A. J., & Sweatt, J. D. (2009). Lasting epigenetic influence on early-life adversity on the BDNF gene. Biological Psychiatry, 65, 760–769. doi: 10.1016/j.biopsych.2008 Roustit, C., Campoy, E., Renahy, E., King, G., Parizot, I., & Chauvin, P. (2011). Family social environment in childhood and self-rated health in young adulthood. BMC Public Health, 11, 949. doi: 10.1186/1471–2458–11–949 Rutter, M., & Sroufe, L. A. (2000). Developmental psychopathology: Concepts and challenges. Development and Psychopathology, 12, 265–296. Sameroff, A. J. (2000). Dialectical processes in developmental psychopathology. In A. J. Sameroff, M. Lewis, & S. M. Miller (Eds.), Handbook of developmental psychopathology (2nd ed., pp. 23–40). New York, NY: Springer.

Sandman, C. A., Davis, E. P., & Glynn, L. M. (2012). Prescient human fetuses thrive. Psychological Science, 23, 93–100. doi: 10.1177/0956797611422073 Sapolsky, R. M., Romero, L. M., & Munck, A. U. (2000). How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews, 21, 55–89. doi: 10.1210/er.21.1.55 Sawyer, S. M., Afifi, R. A., Bearinger, L. H., Blakemore, S., Dick, B., Ezeh, A. C., & Patton, G. C. (2012). Adolescence: A foundation for future health. Lancet, 379, 1630–1640. doi: 10.1016/S0140–6736(12)60072–5 Seeman, T. E., McEwen, B. S., Rowe, J. W., & Singer, B. H. (2001). Allostatic load as a marker of cumulative biological risk: MacArthur studies of successful aging. Proceedings of the National Academy of Sciences, 98, 4770–4775. doi: 10.1073/pnas.081072698 Seeman, T. E., Singer, B. H., Rowe, J. W., Horwitz, R. I., & McEwen, B. S. (1997). Price of adaptation-allostatic load and its health consequences: MacArthur studies of successful aging. Archives of Internal Medicine, 157, 2259–2268. doi: 10.1001/archinte.1997.00440400111013 Segerstrom, S. C., & Miller, G. E. (2004). Psychological stress and the human immune system: A meta-analytic study of 30 years of inquiry. Psychological Bulletin, 130, 601–630. doi: 10.1037/0033–2909.130.4.601 Shalev, I., Moffitt, T. E., Sugden, K., Williams, B., Houts, R. M., Danese, A.,… Caspi, A. (2013). Exposure to violence during childhood is associated with telomere erosion from 5 to 10 years of age: A longitudinal study. Molecular Psychiatry, 18, 576–581. doi: 10.1038/mp.2012.32 Sharp, H., Pickles, A., Meaney, M., Marshall, K., Tibu, F., & Hill, J. (2012). Frequency of infant stroking reported by mothers moderates the effect of prenatal depression on infant behavioural and physiological outcomes. Plos One, 7, e45446. doi: 10.1371/journal.pone.0037385 Shirtcliff, E. A., Allison, A. L., Armstrong, J. M., Slattery, M. J., Kalin, N. H., & Essex, M. (2012). Longitudinal stability and developmental properties of salivary cortisol levels and circadian rhythms from childhood to adolescence. Developmental Psychobiology, 54, 493– 502. doi: 10.1002/dev.20607 Shonkoff, J. P., Boyce, W. T., & McEwen, B. S. (2009). Neuroscience, molecular biology, and the childhood roots of health disparities: Building a new framework for health promotion and disease prevention. Journal of the American Medical Association, 301, 2252–2259. doi: 10.1001/jama.2009.754 Shonkoff, J. P., Garner, A. S., the Committee on Psychosocial Aspects of Child and Family Health, Committee on Early Childhood, Adoption, and Dependent Care, and Section on

Developmental and Behavioral Pediatrics, Siegel, B. S.,… Pascoe, J. (2012). The lifelong effects of early childhood adversity and toxic stress. Pediatrics, 129, e232–e246. doi: 10.1542/peds.2011–2663 Shonkoff, J. P., & Phillips, D. A. (Eds.) (2000). From neurons to neighborhoods. Washington, DC: National Academy Press. Singer, B., Ryff, C. D., & Seeman, T. (2004). Operationalizing allostatic load. In J. Schulkin (Ed.), Allostasis, homeostasis, and the costs of physiological adaptation (pp. 113–149). New York, NY: Cambridge University Press. Singh-Manoux, A., Ferrie, J. E., Chandola, T., & Marmot, M. (2004). Socioeconomic trajectories across the life course and health outcomes in midlife: Evidence for the accumulation hypothesis? International Journal of Epidemiology, 33, 1072–1079. doi: 10.1093/ije/dyh224 Slopen, N., Lewis, T. T., Gruenewald, T. L., Mujahid, M. S., Ryff, C. D., Albert, M. A., & Williams, D. R. (2010). Early life adversity and inflammation in African Americans and Whites in the Midlife in the United States Survey. Psychosomatic Medicine, 72, 694–701. doi: 10.1097/PSY.0b013e3181e9c16f Sternberg, E. M., Chrousos, G. P., Wilder, R. L., & Gold, P. W. (1992). The stress response and the regulation of inflammatory disease. Archives of Internal Medicine, 117, 854–866. doi: 10.7326/0003–4819–117–10–854 Stone, A. A., Schwartz, J. E., Smyth, J., Kirschbaum, C., Cohen, S., Hellhammer, D., & Grossman, S. (2001). Individual differences in the diurnal cycle of salivary free cortisol: A replication of flattened cycles for some individuals. Psychoneuroendocrinology, 26, 295–306. doi: 10.1016/S0306–4530(00)00057–3 Takubo, K., Izumiyama-Shimomura, N., Honma, N., Sawabe, M., Arai, T., Kato, M.,… Nakamura, K. (2002). Telomere lengths are characteristic in each human individual. Experimental Gerontology, 37, 523–531. doi: 10.1016/S0531–5565(01)00218–2 Taylor, S. E., Repetti, R. L., & Seeman, T. (1997). Health psychology: What is an unhealthy environment and how does it get under the skin? Annual Review of Psychology, 48, 411–447. doi: 10.1146/annurev.psych.48.1.411 Taylor, S. E., Lehman, B. J., Kiefe, C. I., & Seeman, T. E. (2006). Relationship of early life stress and psychological functioning to adult C-reactive protein in the Coronary Artery Risk Development in Young Adults Study. Biological Psychiatry, 60, 819–824. doi: 10.1016/j.biopsych.2006.03.016 Taylor, S. E., Way, B. M., & Seeman, T. E. (2011). Early adversity and adult health outcomes. Development and Psychopathology, 23, 939–954. doi: 10.1017/S0954579411000411 Thomas, C., Hyppönen, E., & Power, C. (2008). Obesity and type 2 diabetes risk in midadult

life: The role of childhood adversity. Pediatrics, 121, e1240–e1249. doi: 10.1542/peds.2007– 2403 Troxel, W. M., & Matthews, K. A. (2004). What are the costs of marital conflict and dissolution to children's physical health? Clinical Child and Family Psychology Review, 7, 29–57. doi: 10.1023/B:CCFP.0000020191.73542.b0 Tyrka, A. R., Price, L. H., Kao, H., Porton, B., Marsella, S. A., & Carpenter, L. L. (2010). Childhood maltreatment and telomere shortening: Preliminary support for an effect of early stress on cellular aging. Biological Psychiatry, 67, 531–534. doi: 10.1016/j.biopsych.2009.08.014 Tyrka, A. R., Price, L. H., Marsit, C., Walters, O. C., & Carpenter, L. L. (2012). Childhood adversity and epigenetic modulation of the leukocyte glucocorticoid receptor: Preliminary findings in healthy adults. Plos One, 7, e30148. doi: 10.1371/journal.pone.0030148 United Nations Children's Fund. (April, 2012). Progress for children: A report card on adolescents. New York, NY: UNICEF. U.S. Department of Health and Human Services, National Institutes of Health (2011). 2011 Federal Poverty Guidelines. Washington, DC: Author. Valdes, A. M., Andrew, T., Gardner, J. P. Kimura, M., Oelsner, E., Cherkas, L. F.,… Spector, T. D. (2005). Obesity, cigarette smoking, and telomere length in women. Lancet, 366, 662– 664. doi: 10.1016/S0140–6736(05)66630–5 van den Berg, G. J., Doblhammer, G., & Christensen, K. (2009). Exogenous determinants of early-life conditions, and mortality later in life. Social Science & Medicine, 68, 1591–1598. doi: 10.1016/j.socscimed.2009.02.007 van den Berg, G. J., Doblhammer-Reiter, G., & Christensen, K. (2011). Being born under adverse economic conditions leads to a higher cardiovascular mortality rate later in life: Evidence based on individuals born at different stages of the business cycle. Demography, 48, 507–530. doi: 10.1007/s13524–011–0021–8 van der Vegt, E. J. M., van der Ende, J., Kirschbaum, C., Verhulst, F. C., & Tiemeier, H. (2009). Early neglect and abuse predict diurnal cortisol patterns in adults: A study of international adoptees. Psychoneuroendocrinology, 34, 660–669. doi: 10.1016/j.psyneuen.2008.11.004 Waddington, C. H. (1957). The strategy of genes. London, UK: Allen & Unwin. Weaver, I. C. G., Cervoni, N., Champagne, F. A., D'Alessio, A. C., Sharma, S., Seckl, J. R.,… Meaney, M. J. (2004). Epigenetic programming by maternal behavior. Nature Neuroscience, 7, 847–854. doi: 10.1038/nn1276 Wegman, H. L., & Stetler, C. (2009). A meta-analytic review of the effects of childhood abuse

on medical outcomes in adulthood. Psychosomatic Medicine, 71, 805–812. doi: 10.1097/PSY.0b013e3181bb2b46 Weiner, H. (1992). Perturbing the organism: The biology of stressful experience. Chicago, IL: University of Chicago Press. Whitelaw, E., & Garrick, D. (2006). Epigenetic mechanisms. In P. Gluckman & M. Hanson (Eds.), Developmental origins of health and disease (pp. 62–74). New York, NY: Cambridge University Press. World Health Organization. (2014). Health for the world's adolescents: A second chance in the second decade. Geneva, Switzerland: WHO. Yeh, E. T. H., & Willerson, J. T. (2003). Coming of age of C-reactive protein: Using inflammation markers in cardiology. Circulation, 107, 370–371. doi: 10.1161/01.CIR.0000053731.05365.5A Yehuda, R., Golier, J. A., & Kaufman, S. (2005). Circadian rhythm of salivary cortisol in Holocaust survivors with and without PTSD. American Journal of Psychiatry, 162, 998– 1000. doi: 10.1176/appi.ajp.162.5.998

Chapter 2 Community Violence Exposure and Developmental Psychopathology Patrick H. Tolan INTRODUCTION: VIOLENCE IS COMMON BUT PROBLEMATIC VIOLENCE PERPETRATION AND EXPOSURE ARE IMPORTANT BUT POORLY DEFINED DEVELOPMENTAL PSYCHOPATHOLOGY CONSTRUCTS What Constitutes Violence for Understanding Developmental Psychopathology? Cultural and Societal Variations in What Is Considered Violence Gender and Defining Violence Defining Violence Exposure for Developmental Psychopathology Defining Community Violence and Exposure RATES AND PATTERNS OF EXPOSURE TO COMMUNITY VIOLENCE National Surveys Estimates from Youth Self-Report of Violence Involvement Variations in Exposure by Gender and Ethnicity Age and Exposure Community Violence and Residing in Low- Socioeconomic-Status Urban Neighborhoods WHAT ARE THE MENTAL HEALTH AND BEHAVIORAL EFFECTS OF COMMUNITY VIOLENCE? Meta-Analysis of Effects of Exposure to Community Violence Example Studies Linking Exposure to Psychopathology Studies Relating Other Outcomes to Community Violence Exposure Differentiating or Specifying Effects of Community Violence Exposure RISK AND PROTECTIVE FACTORS ASSOCIATED WITH COMMUNITY VIOLENCE EXPOSURE AND EFFECTS Individual Characteristics Linked to Community Violence Exposure Parenting and Family Characteristics Peer Influences

Neighborhood Social Relations and Structural Characteristics THEORIES ABOUT HOW COMMUNITY VIOLENCE EXPOSURE CAUSES OR INCREASES INDIVIDUAL RISK FOR PSYCHOPATHOLOGY Explanations Emphasizing Trauma Process Social-Cognitive Impact Theories of Community Violence Exposure Effects Stress Models of Community Violence Exposure Effects NEURODEVELOPMENTAL PROCESSES THAT MIGHT BE IMPLICATED IN COMMUNITY VIOLENCE EXPOSURE Genetic Contributions to Violence Exposure Susceptibility Alteration of Neural Networks, Brain Areas, and Regulatory Systems Neurotransmitter Systems Cortisol and Cortisol Regulation Arousal/Resting Heart Rate Implications for Understanding and Affecting Community Violence Exposure INTERVENTIONS FOR COMMUNITY VIOLENCE EXPOSURE Efforts to Reduce Community Violence Efforts to Reduce Effects of Exposure Lessons from School-Based Programs ADVANCING KNOWLEDGE, PRACTICE, AND POLICIES RELATED TO COMMUNITY VIOLENCE EXPOSURE CONCLUSION REFERENCES

Introduction: Violence Is Common but Problematic Violence is a ubiquitous human behavior, occurring across human societies, economic and social positions, and the life span. Yet, for the vast majority of people, direct exposure is rare and is a high impact experience. Similarly, every community has violence, but exposure to violence in the community is exceptional in all but a few locations. Most people are violent at some point in their life and accordingly have some victimization or witnessing experience. Yet even among individuals exhibiting violence at a relatively high rate and for those residing in our most dangerous neighborhoods, violence is still rare during conflict, when facing frustration, or when intending to aggress. Similarly, experiencing violence as a witness or direct victim, while it occurs at alarming rates for some portions of the population, remains in almost each instance an experience that stands out as remarkable. Thus, the experience and

impact of violence exposure, even while having limited direct involvement, has substantial impact on development and warrants careful attention of developmental psychopathologists. This chapter focuses on the impact of exposure to one form of violence, community violence. However, in order to understand the subcategory of community violence, it is important to consider violence more generally and violence in other forms as a human behavior and environmental experience/condition more generally. Evolutionary biology and archeological research explain violent behavior as an adaptive or protective response to compete for resources and power in the social hierarchy and to fend off similar threats from outside the group (Eisner, 2011). Folk and formal literature reveal a longstanding reporting of violence as an undesirable but complex part of human nature dating back to some of the earliest writings (James, 2011). For example, James (2011) notes a record from AD 24 of a trial of Plautius Silvanus for killing his wife by pushing her out a window. His defense was that he was asleep at the time and that she must have leapt. He committed suicide while awaiting trial. Also, his prior wife subsequently was tried for having driven him insane with potions and incantations leading to the murder of his second wife. In this example and in multiple historical accounts from the earliest times, two seminal controversies can be found that persist about the meaning of violence particularly in regard to psychopathology. Is violence a willful criminal act to be treated as a justice issue, or is it an outcome of emotional loss of control or other forms of psychopathology? Is it an inevitable part of human behavior? Are those who commit violent acts psychopathological or simply evil, and if sometimes one and sometimes the other, how do we distinguish these? Is use of violence and harm from it the result of pathological upbringing, or is the developmental course one of less proneness and impact as we mature, such that adults are expected to have full control over that behavior? Is use of violence the result of goading and provocation or a reflection of character and personality makeup? Are there times when violence is appropriate, perhaps even desirable? Is violence exposure inevitably harmful, or does that depend on the type and extent of exposure? What can be done about violence and to help those most vulnerable to its influence? While these considerations fill volumes of legal, philosophical, and social analyses, and violence and related aggressive behavior have received as much attention as any subject in psychology, most of the controversies that are the subject of this scholarly attention remain unresolved. Understanding violence remains of critical interest in understanding psychopathology and is in need of more study. Violence is of continuing interest to those attempting to understand its pernicious role in the developmental course of psychopathology (Cicchetti & Toth, 1995). Violence is implicated regularly in theories about developmental psychopathology, whether as a symptom of some psychopathologies, a putative cause, or the conduit through which psychopathology can be imposed or promoted (e.g., violence causing traumatic anxiety in a witness or victim; Tolan & Gorman-Smith, 2002). In addition, violence occurs in many forms, from a variety of sources, and with differing levels of injury (including death; Tolan, 2007). Capturing the many roles, features, and meanings of violence for developmental psychopathology and organizing them into a coherent explanation can be challenging, because different aspects can seem pertinent depending on the specific interest (Elliott & Tolan, 1999). However, coherently relating the many aspects of violence is essential

in order to understand community violence and violence as a whole and to advance knowledge about its role in psychopathology. This chapter focuses on one form of violence, which is often cited as particularly pernicious: community violence. Still, there is limited understanding of this particular form of violence exposure and its effect on developmental psychopathology. Adequate understanding of the state of knowledge rests on integrating important work from a developmental psychopathology perspective on community violence, on other forms of violence, and on related descriptive and intervention studies. Therefore, this chapter begins with an overview of the multiple definitional challenges in determining what violence is, in general, and how it relates to the common view and framework of community violence. This general-to-specific relation is pertinent in the ensuing sections on patterns of violence exposure, causes of exposure, effect of exposure, interventions and policy related actions, and suggested further research and actions.

Violence Perpetration and Exposure Are Important but Poorly Defined Developmental Psychopathology Constructs Because almost all people are violent at some point in their life and almost all are exposed to some form of direct violence, it is noteworthy that violence nevertheless is seen as dramatic and/or problematic in all but a few instances (e.g., minor physical aggression among young children, righteous war, self-defense, fighting injustice). Also, the harm is not seen as limited only to direct exposure or to whether that exposure is experienced as victim or perpetrator. Indeed, there is little disagreement that violence is almost always problematic and that abating it is positive. However, beyond that matter, consensus drops quickly about how it is understood and what to do about it. Much of the disagreement is tied to how violence and violence exposure are defined. In addition to carrying differential assumptions and connotations, the definitional variations make scientific comparison and certainty difficult. Moving to a useful scientific understanding of violence requires a balance between precision and consistency in definition and a consideration of the differential interests in research, practice, and policy related to violence exposure. That tension needs identification and articulation to facilitate sustained shared meaning and related action.

What Constitutes Violence for Understanding Developmental Psychopathology? Among the controversies related to violence are these questions: What is considered violence? How is violence differentiated from nonviolent aggression and other forms of interpersonal disregard or insult (Elliott & Tolan, 1999; Tolan, 2007)? Variations in these boundaries carry important implications for the role of violence in psychopathology, how patterns are identified, what and how risk factors are related, and which interventions and policies seem appropriate. Moreover, different definitions can create variations in data, testimony, and other information that are the bases for scientific and public understanding, variations that in turn can lead to

divergence in implications (Loseke, Gelles, & Cavanaugh, 2005). The lack of consensus hinders scientific progress by obstructing coordination and comparison between epidemiological and risk/causal studies, obscuring the meaning of trials of program and service procedures effects, and undermining standards of judging effectiveness. A lack of a consistent shared definition of violence impedes scientific interests such as generalization, replication, and robustness of knowledge. This vagueness of definition also degrades the respect given to such evidence in comparison to opinions and how policies designed to address violence define expertise. Common and Differentiating Features of Violence The common feature of most violence definitions is intent to injure from a perpetrator's point of view and physical harm or fear from threat of such harm from a victim's point of view (Krug, Dahlberg, Mercy, Zwi, & Lozano, 2002). Forms of violence are distinguished along dimensions that also form legal distinctions about violence: intentionality of action and a level of notable harm. This definition differentiates violence from injury that may result from an accident (no intention), negligence (failure to show due caution that results in harm), and recklessness (acting in such a manner as to greatly increase the potential for injury), not by the harm that may result by criminial intention (mens rea). Mens rea is legally an essential component of crime perpetration. This definitional base also emphasizes physical harm or the threat of such harm (e.g., the threat to hit if someone does not do as the perpetrator insists) as an essential determinant that violence has occurred. Violence can be narrowly defined as an act of intentional physical aggression inflicted in order to cause and resulting in physical injury of self or another. Constraining the definition in this way has the advantage of making determination of what is violence easy and tracking violence patterns and effects more straightforward. Such clarity may enable criminal and civil codes to guide legal procedures more easily. It also may be more important for tracing epidemiological patterns than including less certain acts and intentions (Tolan, 2007). However, these benefits may come at the cost of excluding many important nuances of violence experience. Such a narrow definition simplifies three components that vary across violence definitions: (1) role of intent, (2) what constitutes an aggressive act, and (3) how harm to the victim(s) should be considered. For example, violent intentions may not result in injury. Recklessness, while a less intentional condition leading to harm, can be as injurious as other forms of violence and can appear intentional to the victim. Similarly, violence may not always be expressed directly or experienced directly. It can come in the form of intimidating gestures or verbal threats of violence. Participants in a conflict may have quite different perspectives on whether violence was present, using different thresholds of what is termed physical threat and what needs to be explicit or can be implied to constitute violence. For example, a group of young men on a city corner may see as a threat another group of young men from a different neighborhood moving physically closer to them. They may be correct if the groups are organized gangs competing for drug sales turf. They also may not see such a move as violence, as no one was actually attacked or hurt, or no direct threats were made. Similarly, a couple bickering loudly within earshot of neighbors may induce fear and

worry about the female's safety for some, but others may see the encounter as dramatic but insubstantial. The arguing couple may each experience the conflict as imminently at risk for violence or not. Are Physical Harm Threat and/or Intent Necessary for an Act to Be Violent? There is less certainty, and substantial disagreement, about how fully intentional the expression to cause physical harm must be for the act to be considered violent. The extent to which intent is seen as required for violence to occur is important for understanding how violence arises in relation to psychopathology. Similarly, the likelihood or imminence of intentional harm can affect how violence exposure is tied to trauma. It may be that traumatic impact is significantly dependent on actual harm (compared to threat or witnessing violence) or it may not be. Similarly, it is an unsettled debate whether, for violence to be present, the intention must be to cause physical harm or merely to coerce another (Tolan, 2007). For example, most would agree that threatening to hit someone unless he or she does as you demand is violent. However, in a given study, there may be important variations in impact and basis for effects depending on whether threat is treated the same as physical injury and how matters such as perceived intent and imminence are considered. At present, most of the difference in inclusion and exclusion across studies and reviews is not based on careful definitional comparison or empirical tests of the role of intention, actual harm versus fear of harm, and/or imminence of harm. Instead, variation in measurement makeup often is overlooked. Also, frequently a measure shows that a particular perspective was applied without the basis of the theory being clearly acknowledged (Chalk & King, 1998). Stakeholders vary widely on where such boundaries should be drawn (Jackman, 2002; Jouriles, McDonald, Norwood, & Ezell, 2001; World Health Organization [WHO] Summit on Violence, 2002). For example, while some definitions restrict violence to imminent threat of harm, others expand it to include direct attempt to oppress or coerce, and others go further, including where it might be implicit or would be considered so by an outside observer. These boundaries have been debated most when violence is occurring in a relationship and there is differing power and ability to cause harm (e.g., domestic violence, child abuse). In regard to a recipient or victim's perception of potential harm or threat of injury as part of the definition of violence, some argue that acts, orientations, or statements that are experienced as intimidation, coercion, oppression, or creating undue insecurity are violent, even if they do not involve actual physical aggression or specific verbal threats. From this viewpoint, if the “victim” feels threatened or afraid or identifies some act as violent, it is considered violent. Although this definition incorporates the experiencing of violence and informs about behavior during conflict, it also introduces unspecified variation in responses within samples and across studies. Moreover, we cannot see any differences among the contribution of violence, the perspective about violence, and personal reaction when they are all collapsed into defining any interaction as violent. Others suggest that violence should be differentiated from the victim's perception of threat even if only to permit more careful empirical testing of the relation between acts and perceptions (Johnson, 1995; Malley-Morrison & Hines, 2004). For example, in such an approach, the act

would be defined as violent by intention to injure or to directly intimidate through threat of violence. Typically, some determination is made about what is a reasonable or common basis for concluding that threat has occurred (akin to the legal criterion of “reasonable person standard” for perception and expectation). This is done so that variations in labeling and reaction can be studied and understood as important but distinct components. When defined with objective criteria rather than personal determination, it can be argued that more sensitive consideration of variations in what is threatening also can be better considered because the perception and the conventional/objective definitions can both be considered. In addition, proponents of such a distinction would note that it also helps in offering perspective in explaining violence and the effects of the relationship between the involved persons, such as ongoing high levels of interpersonal conflict, contemptuous attitudes, or neglect of expected care. It is argued that limiting the definition of violence to the act itself and by using a “reasonable person” standard of threat/harm, research can disentangle violence and its affects from co-occurring problems. Research also can differentiate between related but distinct contributors to functioning, susceptibility to trauma or other pathology effects, and best interventions.

Cultural and Societal Variations in What Is Considered Violence Another important challenge in defining violence is that cultural differences may affect the meaning of the terms violence and injury (Hudley & Taylor, 2006; Walters & Parke, 1964). For example, injury in some cultures is broadly defined as an attempt to harm or manipulate the well-being of others. Alternatively, in other cultures, injury is reserved for serious physical harm, and minor assaults may not be considered violence. Similarly, behavior that is considered very offensive in one culture may be considered acceptable, even expected, in another. Even if physical force is clearly evident, such acts may not be seen as violent, or they may not be treated as similar to other acts of violence. For example, if two friends are teasing each other in a display of relative dominance during a group gathering of adolescents, one might smack the other upside the head. This action would not be seen as violence by anyone familiar with the exchange; it would be seen as part of group hierarchy competition in certain contexts. If asked about violence exposure, neither the perpetrator nor the victim would report this event. Outside observers might say that this exchange was clearly violent, even if the adolescents involved did not view it that way. Whether cultural acceptability and commonality of occurrence should be considered criteria for violence remains controversial (Tolan, Dodge, & Rutter, 2013). For example, in a state of armed struggle, exposing children to violence may be seen as unavoidable. Therefore, those responsible for their safety and well-being may focus on how that exposure is managed and how the child might learn to avoid such exposure. Similarly, outsiders may view increasing empathy for an enemy and trying to train youth not to act violently toward members of the enemy's group as virtuous. However, within the culture, such actions may be seen as confusing and harmful to the children. Turning the other cheek may diminish children's vigilance, which risks their safety and/or distracts from allegiance to a religious cause or patriotic interest (Garbarino, 1996). A more common example of violence

as part of developmental psychopathology is the question of how violence by parents affects youth when it is a part of routine discipline (Lansford et al., 2005). Although studies provide mixed results, a trend shows that cultural variation in harm of corporal punishment is related to how typical such practice is and how acceptable or normative it is perceived to be. It may be that acceptance of minor forms of punishment vary by culture whereas severe physical punishment may not be. The lesser forms of punishment also may be part of what is viewed as parental responsibility (or hierarchically, by teachers to students or older children to younger children). Cultural considerations in the definition of violence also apply to understanding community violence, particularly how fear and perceived safety are related to actual levels of harm. For example, in the United States, there is a belief that violence in schools is increasing in the form of sudden violent rampages resulting in multiple victims. Each such horrific incident receives wide attention in the media, but it is important to recognize that such incidents still are extremely rare. At the same time, other types of school violence, which in fact causes more victims per year, do not receive such attention. These acts include student-to-student, studenton-teacher, and teacher-to-student violence. This violence often is concentrated in schools that serve lower-socioeconomic-status (SES) and ethnic minority populations. And, while some schools account for a higher percentage of overall violence, most schools are safe and in fact remain among the safest environments for children across neighborhoods. These variations in what affects the perception of school safety means that actual and believed threat may diverge, to the point that concern could be diverted to circumstances where violence victimization is actually unlikely (Tolan, 2001a). Similarly, fear of strangers abducting children led to an infusion of funding and programming to teach children about “stranger danger,” even though such occurrences are much less common than abduction by a person known to the child and is much less likely than harm from family members. Such emotionally charged responses seem to misdirect from acknowledging violent threats by family and acquaintances in favor of a much less likely threat of violence from strangers. For example, it may be presumed that violence exposure is not an issue in middleand upper-middle-income communities because such communities have lower nonfamily assault and homicide rates than poorer and urban counterparts. Yet victimization, witnessing, and awareness of violence occurs within such communities. Conversely, it does not accurately characterize the daily experience of most urban, lower-SES residents (Horney, Tolan, & Weisbourd, 2012). These examples indicate that the perception of a setting as violence prone or violence ridden can induce fear, hypervigilance and mistrust, attitudes that affect the definition of violence and how it may affect development.

Gender and Defining Violence Related to the cultural and societal variation in violence perceptions is the understanding of the role that gender plays. Evidence clearly shows that males perpetrate violence at greater rates and are exposed to greater levels of violence than females (Farrington, Langan, & Tonry, 2004). Gender differences, particularly in physical aggression, are present early in life and remain throughout development (Tremblay et al., 2005). Further, it is widely recognized that

male aggression includes more violence, ability to harm, intimidation, and other threatening aspects than female aggression. Also, community violence disproportionately occurs by and to males, although the extent of difference is hard to determine from existing research. Yet among community samples of male-female couples, overall rates of violence do not differ by gender, and most violence is rare and minor (Tolan, Gorman-Smith, & Henry, 2006). A key consideration is that male relationship violence is more likely to be part of a pattern of coercion, intimidation, or contempt. This difference in relationship power is considered important in defining violence and in locating concerns about battering, political and economic inequities, and social resources when characterizing gender-based violence (American Psychological Association [APA] Presidential Task Force on Violence and the Family, 1996; Jouriles et al., 2001). Whether attempting to incorporate status differences in defining violence overtly or noting this contextual determinant of how violence and threat of violence might be related to gender, it is evident that the relation of exposure to psychopathology is dependent at least in part on this variation. These considerations have been central in research and policy debate about partner violence but have received much less attention in the study of community violence. In part, this is because much of the interest in violence emanates from study and policy related to domestic violence. Also, domestic violence is differentiated from what is considered community violence, treated as different in regard to likely effects on witnessing and with different treatment legally (likelihood of legal intervention and seriousness of potential charges). Whether status difference is literally less pertinent for community violence or not, partner violence is overheard regularly by neighbors. Thus, partner violence can also be community violence. It spills into the community experience, bringing to light gender and related power and threat differences for measuring and evaluating community violence exposure. Moreover, there is controversy about how cultural variations and how gender should be incorporated into violence definitions. For example, even if actions that are violent, if violence is defined as physical harm or threat of such harm, are legal or sanctioned, should they be treated as violence in studies identifying patterns and rates and relating outcomes to risk factors? Should acts that seem to diminish the rights or status of others be treated as violence, even if they are labeled “not violent” in a given culture (Fagan & Browne, 1994)? When an act seriously harms the viability and safety level of a community, even if the act is legal or sanctioned, should it still be considered violent? Is the failure to care for those in pain or the imposition of prolonged neglect an act of violence? Not surprisingly, some argue for culturally based definitions of violence, whereas others argue for absolute definitions that focus on physical harm or imminent threat of harm. Some believe variations should be measured consistently and then interpreted within cultural contexts and other potential influences (Farrington et al., 2004; Krug et al., 2002). One argument for such an approach is that the relation of gender to violence and other cultural and societal variations that affect how a given act is perceived, experienced, and viewed by others needs to be considered. The best way to do that is to differentiate these variations from what is called violence. A second argument is that these effects are marked through legal codification, traditions, social structures and other engrained beliefs, so translation of scientific findings using this differentiation is more in line

with social policies and existing norms. A third argument is that these considerations are always present, albeit in varied importance and prominence, and as such they should be measured separately for full consideration even if they are not implicated in a specific experience of a given individual (Chalk & King, 1998; Tolan et al., 2006). These arguments seem to permit due accounting of the role of gender in violence without conflating gender with violence.

Defining Violence Exposure for Developmental Psychopathology These variations in what is considered violence and how it is measured affect studies of violence exposure and its impact, yet specifics of measurement formulation are rarely described in empirical studies. While giving due consideration to these variations in perspective, concerns, and empirical basis, there is consistency across studies that points to the experience or imminent threat of physical harm as the hallmark of violence. This does not mean other forms of aggression and threat are not harmful or can be neglected, or that corollaries and corresponding problems are not worthy of consideration in understanding violence as a developmental psychopathology issue. However, focus on this central distinguishing feature seems important to permit differentiation and organization of the relation of violence exposure to closely related experiences and to conditions, covariates, and group differences in effects. Thus, there is increasing agreement that violence should be limited to referring to intentional physical harm or immediate threat of such harm to coerce another. Even these basic considerations could lead to definitions that may vary in what is considered violence exposure. Because violence is a politically, socially, and culturally determined construct, there is an unavoidable compromise required between interests such as definitional precision and replication comparability and encompassing varying applications for policy and other actions (APA Presidential Task Force on Violence, 1996; Chalk & King, 1998; Office of the Surgeon General, 2001; WHO, 2002). For purposes of this chapter, the definition rendered by the World Report on Violence and Health (WRVH; WHO Summit on Violence, 2002, p. 5) seems to capture the variations in violence forms and sources while distinguishing it from other related behaviors. It is also suggested here as a definition for future work in the field: the intentional use of physical force or power, threatened or actual, against oneself, another person, or against a group or community, that either results in or has a high likelihood of resulting in injury, death, psychological harm, maldevelopment, or deprivation. The WRVH Typology of Violence The WRVH also provided a typology of violence and categorized it into four modes of injury: physical, sexual, psychological, and deprivation. As is evident from this differentiation, there is interest in discerning when the injury is through sexual violence. The categorization also shows that in addition to physical injury, there can be injury that is psychological or that occurs by depriving the victim of basic needs or safety. These details help to define more specifically the forms of harm or likelihood of harm that are important in understanding violence and

differentiating it from other potential risk influences. WRVH also differentiated violence into three subtypes based on the victim-perpetrator relationship. Collective violence refers to violence committed by large groups of individuals, connoting the role of the group in precipitating the violence of the individual members. Collective violence includes acts that can be social group, political, or economically based. Self-directed violence refers to violence in which the perpetrator and the victim are the same individual, with distinction between self-abuse and suicide. Interpersonal violence refers to violence between individuals, and can be further differentiated into family/intimate partner violence and community violence. The former category includes child maltreatment, intimate partner violence, and elder abuse, while the latter is differentiated into acquaintance and stranger violence. With these categorizations and differentiation of forms of harm, this WHO definition brings into relief the complex variations that violence exposure can take and some systematization of how to measure and compare variation impact by type and relationship of victim and perpetrator. By these categorizations and in most writing, community violence includes: youth violence; assault by strangers; violence related to property crimes; and violence in schools, workplaces, and other institutions. Considering Neglect or Deprivation as Violence The word power, sometimes used in defining violence perpetration, refers to use of power in relationships and in relative social standing as a basis for imposing physical harm, threat of loss of safety or essential care, or deprival of resources (neglect). This view of coercive use of power as violence has focused most on relating gender to understanding violence impact. The inclusion of the use of power also incorporates treating acts of omission or deprivation/neglect as being violent. In order to understand violence as part of developmental psychopathology, this inclusion is important; it allows consideration of how coercion and harm can arise from lack of expected care or concern. However, this inclusion seems to suppose that effects of deprivation of care or regard are similar in extent and process as physical threat or harm rather than being based in empirical support. This relation is far from clear. In part, this is because many samples in studies to identify effects on maltreatment included children exposed to neglect as well as those physically or sexually abused. In most such cases, the large majority of such children have been neglected as well. These two limitations prevent understanding of how effects of neglect relate to other forms of victimization (Trickett & McBride-Chang, 1995). However, when differentiation permits comparison, results are mixed about whether neglect and violence have the same pathology-inducing and function-impairing effects (Trickett & McBride-Chang, 1995; Twardosz & Lutzker, 2010). The way in which neglect relates to actual physical harm is relevant to community violence because community violence often includes exposure that is indirect or that induces fear. From that view, the absence of a safe environment is implicated as one form of violence exposure. For example, does failure to provide children with a safe route to school constitute the same harm as witnessing violence during that route? This WRVH definition also includes a broad range of violence effects that stretch beyond

variations in physical harm to include an increase in frequency of psychological disorders, elevated susceptibility to such disorders, and failures in development. While not stated explicitly, the definition also permits considerations in alteration of developmental pathways and sensitivities that can evince diminished functioning at key transitions or specific situations and at other points in the life span. Thus, consequences can be immediate as well as latent and can last for years after the initial violence. Also, it is important to consider not only the effects of violence on victims but also to examine the effects on perpetrators, those emotionally connected to the victim or perpetrator, and those witnessing the violence either directly or indirectly (e.g., child experience of partner violence between parents).

Defining Community Violence and Exposure If violence is defined as imposing intentional physical harm or as the imminent threat of such harm, then community violence is the form of these actions that occur in one's community. Community violence is often differentiated from other violence in that it is something other than family or partner violence or violence between friends or acquaintances. It is also differentiated by where it occurs (or actually by where it does not occur). Community violence usually is defined as exclusive of violence in the home or at work. Violence at school is sometimes differentiated from community violence but most often is treated as a subset of community violence or separate therefrom. A third differentiating characteristic of community violence is that it includes a broader set of experiences than most other forms of violence. It is thought to include not only direct injury and threat of such injury but also knowledge or witnessing it. Often simply residing in a neighborhood with elevated rates of crime and violence is considered to be exposure to violence—growing up in an unsafe neighborhood with elevated opportunities to witness violence or know persons affected by violence can potentially affect development. Community violence is differentiated from family violence by the relationship between the perpetrator and the victim or witness. Also, as family violence means violence between persons who are also often responsible for care and affection toward each other or the child, the emotional and relationship complexities are differentiating. Similarly, serious injury or death of a family member caused by someone other than the family is usually distinguished from community violence. It should be noted that while these definitional differentiations are increasingly considered in measuring impact of different forms of violence, measurement of community violence varies considerably from study to study in what is included, how well it is differentiated from other forms of violence, and to what extent community violence is considered along with other forms (often only one form is considered in any given study). Therefore, community violence exposure refers to a broad range of experiences related to violence and should be considered on a study-to-study basis. For example, violence exposure may refer to living in a high-crime community in one study, or it may be limited to serious violence that community members have heard about. Or it may be measured as personal experience of being the victim of physical aggression or witnessing it. Because community violence is often of interest as much for its direct effects as for its impact on creating an unsafe environment, many studies collapse varied contact with violence, such as victimizations of

friends and families, witnessing or learning/knowing of violence in the surrounding community, with firsthand victimization. Thus, to some extent, community violence can refer to some subset or a variety of forms of violence that do not occur between family and friends and are not the result of war or political actions. Also, a given study may not distinguish, in measurement or in calculation of variable relations, the differences among direct victimization from exposure due to being a perpetrator of violence from witnessing violence as a bystander, from merely knowing about recent violent events, or a perception of low safety personally or of the school overall (Tolan, 2001a). Differentiating Types of Community Violence Exposure There is increasing agreement regarding the need to measure each of these types of exposure and to test for differentiation of effects (Tolan, 2001a). For example, Shields, Nadasen, and Pierce (2009) reported that direct victimization had more substantial and sustained effects on children's social and psychological distress than witnessing or hearing about violence, but the latter two forms were not significantly differentiated in regard to effects. However, the pattern of findings suggests it is not merely a matter of extent of exposure or proximity to victimization that accounts for impact. Along that line, Schwartz and Proctor (2000) found that violent victimization was associated with negative social outcomes whereas witnessing violence was related only to subsequent aggressive behavior. Moreover, victimization effects on outcomes were mediated through emotion dysregulation. In a study with a related but different focus, Gorman-Smith, Tolan, Sheidow, and Henry (2002) tracked the link between family violence involvement and community violence and found that it was rare that community violence perpetration occurred without concurrent or preceding violence at home. Yet family violence did not simply account for effects of violent behavior in the community. Empirically tracing the role of violence exposure to youth development and psychopathology patterns, causes, and course can entail extensive complexity. In fact, measurement detailed enough to make all the differentiations noted here as potentially relevant is impractical even for large studies with considerable resources. Also, as initially noted, it is unlikely for a given person to engage in violence even if subjected to it over substantial amount of time. Serious incidents of violence are rare enough to strain statistical reliability and validity assumptions for accurate interpretation of patterns even in high-crime communities. For these and other reasons most of the literature has focused on violence exposure overall, with a few instances in which differentiation of community violence from family violence is incorporated. What is known about demographic patterns, relation of forms of exposure, correlates, risks and protective factors, and effective interventions is often about exposure as generally defined or if there is more differentiation of type, role in the violence, or extent of exposure it is specific to the interest of the particular research project. There is little consistency across studies in such differentiation.

Rates and Patterns of Exposure to Community Violence Exposure to community violence came to the fore as a public health issue because surveys

evidenced high rates of violence victimization and witnessing among youth residing in loweconomic inner-city communities (Gorman-Smith & Tolan, 1998). These reports coincided with empirical verification of traumatic stress in veterans due to exposure to violence of war conditions and independent reports of posttraumatic stress symptoms, similar to what was seen in war veterans, among children exposed to terrible events. These events, such as kidnapping, produced interest in the conditions of a community, particularly violence levels, might potentially harm children (Tolan & Guerra, 1998). A third coincident influence was the emerging literature documenting effects of child abuse and neglect on developmental psychopathology (Cicchetti & Toth, 1995, 2005). As with traumatic impact of war and disaster, the finding of links between abuse and psychopathology susceptibility and developmental course helped to persuade many scientists and policy makers that violence in the community could have direct effects on child development. Many of the initial reports about community violence rates, while useful in marking such violence a risk factor of importance, were dependent on questions that often were general, did not make clear the time frame for response, and did not provide ways for children to differentiate types of exposure. Moreover, there was considerable reliance on convenience samples via one-time surveys and limited attention to population representation and variations (Margolin & Gordis, 2004). Only a few of these studies differentiated questions about community violence exposure from familial or acquaintance violence and that which was a byproduct of perpetration of violence from exposure due to victimization (Gorman-Smith & Tolan, 1998). Nevertheless, what was shown was that some children, particularly those residing in impoverished inner-city communities, experienced community violence at alarming levels, with reports as high as 70% reporting violence exposure (Thompson & Massat, 2005). For example, several studies published in the early to mid-1990s estimated that about 25% of inner-city youth have witnessed someone shot and/or killed during their lifetime (Groves, Zuckerman, Marans, & Cohen, 1993). The shock of one in four children witnessing such serious violence as well as documentation of related psychopathology led to much interest in rates of violence exposure of youth in the United States. As with child deaths due to violence, the United States stood out from other countries in the rates of children exposed to serious violence (Tolan, 2001b).

National Surveys More careful assessment and sampling, although they documented for lower rates overall, seemed to verify the exceptional level of community violence exposure of U.S. youth. It also verified elevated rate of exposure in economic poor urban (inner-city) communities. For example, a recent national sample of U.S. adolescents found that 55% reported some type of violence exposure (McCart et al., 2007). Similarly, another national sample study estimated that 17.6% of children and adolescents are direct victims of an assault in the community (Finkelhor, Turner, Ormrod, Hamby, & Kracke, 2009). One of the most extensive surveys has been by Finkelhor and colleagues (Finkelhor et al., 2009; Finkelhor, Turner, Shattuck, & Hamby, 2013), who reported on exposure to multiple sources of violence. Utilizing an extensive questionnaire, the Juvenile Victimization Survey, on a national sample of children

ages 2 to 17 (adult reported for those under 10), found that 33% reported witnessing violence in the past 12 months, with 69% reporting some exposure to violence during their lifetime. In another report from the same data, Finkelhor et al. (2013) found that 22% of children had witnessed violence in their homes and schools in the past year and 39% had witnessed violence against another person during their lifetime. Specifically, 16.9% reported witnessing an assault of another in the community in the past year, with a lifetime prevalence of 27.5%, and 3.4% reported indirect exposure to community violence in the past year, with a lifetime prevalence of 10.1%. Notably, these rates are higher than for witnessing violence toward a family member or in the home. Other studies that have differentiated community violence from other exposure led to national estimates of witnessing community violence that range from 19.2% to 37.8%, with rates in inner-city communities of 30% to 50% directly experiencing violence and over 70% witnessing some form of serious violence (Finkelhor et al., 2009; Stein, Jaycox, Kataoka, Rhodes, & Vestal, 2003; Tolan, Gorman-Smith, & Henry, 2004;). Among the major finding from this survey was the pattern of polyvictimization—that is, children exposed to violence were often exposed to more than one form. As the number of forms of violence they were exposed to increased, so did reported distress and problems in functioning. The results may suggest that the harm that results from multiple exposures does not increase linearly as exposure increases but may have qualitative increases in harm. However, this study leaves undifferentiated the extent to which effects from exposure overall differs from overall and component types of exposure. Repeated exposure and exposure to multiple types of violence are rare but are strongly related. In addition, multiple exposures may be attributable to additional risk factors also related to psychopathology. Thus, while documenting that multiple exposures occur and may have a nonlinear impact on functioning, this correlational survey study cannot authoritatively provide understanding about how exposure across types might impact psychopathology. The Centers for Disease Control (CDC) has used the Youth-Risk Surveillance Survey to track exposure to and engagement in violence in school and the community. Although it does not differentiate fully community versus school or in-home/familial violence, the questions utilized permit reasonable estimates of community violence exposure. The most recent report, for 2013, indicated that nationwide, 24.7.8% of students had been in a physical fight one or more times during the 12 months before the survey, with the rate higher among males (30.2%) than females (19.2%) and higher rates by ethnicity (CDC, 2014). Blacks had a higher rate (34.7%) than Whites (20.9%) and Hispanic youth (28.4%). Also, Whites and Hispanics differed significantly. Within gender, the ethnic differences were maintained, but within ethnicity, the rate difference for males to females was similar. As with many violence exposure indicators, the rate has decreased since 1991 from 42.5% to 24.7%, with a significant drop between 2011 and 2013 (32.8%). The number of youth who reported having been threatened or injured with a weapon on school property one or more times during the preceding 12 months was 6.9%; 8.1% of youth reported being in a physical fight at school. Each indicator followed the same pattern of higher rates among Black than White or Hispanic youth as well as higher among males than females. In most cases, the differences between Hispanic and White youth were significant as well. The same ethnic group pattern was found for forced sexual intercourse (7.3% overall),

but women reported rates of about twice to three times what males reported (10.5% to 4.2% overall respectively). Also, unlike most other physical attacks, significant differences in rates between 2001 and 2103 were found. Being bullied at school (the only setting asked about) showed a different ethnic and gender pattern. Overall, this was reported by 19.6% of students with females reporting higher rates than males (23.7% versus 15.6%) and a higher rate among non-Hispanic White students than Black or Latino students. Similarly, with an overall rate of 7.1% indicating they had not attended school at least 1 day during the 30 days before the survey due to feeling unsafe at school, females had a higher rate than males. This rate was higher among Blacks than Hispanics. Both rates were higher than for non-Hispanic white students. Notably, the rates of such behavior increased from 1993 to 2013 rather than decreased, as has experience of violence at school and overall. Across indicators, irrespective of gender or ethnicity trends, there was considerable variation in exposure by state and among similar communities. For example, across 37 states, the prevalence of having been in a physical fight ranged from 16.7% to 31.0% with a median of 22.8%. Across 19 large urban school districts in the survey, the median prevalence rate was 26.3% with rates ranging from 17.2% to 37.6%. Although not summarized here, comparisons do tend to find urban-suburban differences in violence exposure rates.

Estimates from Youth Self-Report of Violence Involvement Another way to estimate exposure to community violence, particularly of youth, is to consider how youth report weapon carrying, which suggests either likelihood of violence or concern for victimization. Notably, the vast majority of weapon-carrying youth say they do so for protection against attack rather than as intention to harm others (Tolan, 2001b). The CDC Youth-Risk Behavior Study (2014) reported 17% of youth had carried a weapon at least 1 day in the past 30 days, (more than 1 in 4 males; only 1 in 12 females). The highest rate was for White males (33.4%) followed by Black Hispanic males (23.8%) and Black males (18.2%). Female rates did not differ significantly by ethnicity. Gun-carrying patterns mimicked the overall weapon-carrying rate (5.5% overall; the highest rate, 10.7%, was for White males). There was no significant difference between Black and White males (Black males 9.8%, White males 10.7%)]).

Variations in Exposure by Gender and Ethnicity These patterns reveal consistency in greater exposure for males than females for most types of violence, particularly community violence. Similarly, minority ethnic groups tend to have higher rates of community violence exposure, a fact that seems to be closely related to differences in likelihood of residing in a community marked by concentrated poverty and other associated problems, including higher violence rates. Therefore, it seems that differences in exposure rates are mostly accounted for by community economic levels and related resources. However, these covarying neighborhood conditions also represent a complex pattern of elevated risk, lesser opportunities, and scarce resources for avoiding exposure and for

managing exposure (Tolan, Guerra, & Montaini-Klovdahl, 1997). These conditions constitute a differential risk ecology that may well relate to differential bases for exposure, impact of exposure, and ability to mitigate exposure and impact. The gender differentiation seems to cut across ethnic groups and economic levels (to the extent studied). In this case, the greater exposure does not correlate with the group that has greater social status or access to more resources for coping. In addition, as with other forms of violence, serious community violence is more likely to emanate from males than from females. Thus, the differential exposure seems to reflect differences in proclivity for involvement in settings and antisocial behavior that is related to violence. Notably, the few studies that have made gender comparisons show consistency in relation to symptoms and impact on functioning, although exposure rates tend to be higher for males (Fowler, Tompsett, Braciszewski, JacquesTiura, & Baltes, 2009; Pinchevsky, Wright, & Fagan, 2013). Yet in one of the most thorough comparisons of forms of exposure, Pinchevsky et al. (2013) reported that direct (victimization) and indirect (witnessing or hearing about) community violence proved to have limited relation to substance use for males (alcohol or marijuana use) but significant relation to use for females. The effect was more robust for indirect exposure than direct exposure. Notably, all of the relations were small, and many relations were not significant.

Age and Exposure Exposure to violence in general and to community violence in particular is usually related to age, with adolescence marking a time of substantial increase in both time spent in the community and unsupervised time there. This means that youth not only have greater opportunity for exposure to violence, but violence also may occur in part because adult supervision that could quell it is not present (Tolan & Guerra, 1998). The exception to this general trend is for physical assault when violence (abuse) from family members is not differentiated. However, Finkelhor et al. (2013) state that across maltreatment sources, rates of maltreatment reported for children ages 2 to 5 was 10% ; ages 10 to 13, 17%; and ages 14 to 17, 21%. Physical assault rates did not increase over time. Witnessing violence increased from a rate of 8% of those ages 9 and under, to 26% for those ages 10 to 13 and 43% for those ages 14 to 17 (Finkelhor et al., 2009). Indirect exposure increased from a rate of 2% for ages 9 or younger to 4% for ages 10 to 13 and 6% for ages 14 to 17. Another trend evident across studies is that severity of violence exposure increases with age. For example, in the Finkelhor et al. survey (2013), those under 9 years of age were most commonly exposed to assault without a weapon or without injury and assaults by a sibling, with 10- to 13-year-olds reporting more assault with a weapon, sexual harassment, and kidnapping and reporting more witnessing of domestic violence in the home involving their parents and assaults by other family members. Fourteen- to 17-year-olds had the highest rates of any age group for the most serious forms of violence victimization. Among this age group, the most common types of exposure were injurious assaults, gang violence, sexual assaults, physical and emotional abuse, and witnessing violence in the community. They also reported more sexual violence victimization than younger children.

Community Violence and Residing in Low- Socioeconomic-Status Urban Neighborhoods Implicit in results of the CDC surveys and others is that exposure to community violence is related to residing in urban neighborhoods with concentrated poverty, or what has been termed the inner city (J. Wilson, 1987). These neighborhoods are marked not just by elevated rates of violence, but low home ownership, the absence of middle-income residents, and many other social and political problems. There is now considerable evidence that living in such neighborhoods is associated with exposure to a host of risk factors, reduced protective factors, and many other poor outcomes for children, including aggression, violence, substance use, depression, and academic failure (Gorman-Smith et al., 2002; Leventhal & Brooks-Gunn, 2000; Sampson, 1997). Surveys of children and adults reporting about children sampled from lower-SES urban neighborhoods routinely report high levels of exposure (40%–70%) to the most serious forms of community violence, setting these neighborhoods apart from most other U.S., communities regarding child risk for exposure to community violence (Tolan & GormanSmith, 2002). The juxtaposition of patterns evident in the CDC survey and others—which show greater rates among those residing in low-SES, urban neighborhoods—seems to explain much of the difference in exposure for minorities, who reside disproportionately in these neighborhoods. It is not clear how much of the difference is due to the compilation of risk factors, such as urban residence and lower economic status and the concomitant dearth of resources, or how much is the confluence or coincidence of these factors for given neighborhoods. However, most studies to date have not systematically compared what might differentiate exposure across such communities or compared them with communities that differ on one but not other risk characteristics. Most of the literature focused on community violence has used samples from communities that are urban and have low socioeconomic levels. As such, this population has comprised a large proportion of minority ethnicity samples, but without due recognition that these characteristics might differentially relate to violence exposure and impact. Nor is adequate consideration given to how the conflating of these two demographic characteristics may lead to unjustified conclusion that it is both ethnicity and low income prevalence that promotes community violence. Because the majority of studies also focus on the variation of exposure for individuals within high-risk communities, the focus tends to be on how individual and family variables explain exposure and impact rather than on how community differences and demographic differences might explain them. Thus, the understanding based on current findings is constrained to documenting that rates drawn from such neighborhoods are high and that relations to psychopathology can be found in many studies, some of which are focused on youth from urban, poor communities. For example, in a study of youth receiving mental health services, community violence, domestic violence exposure, and maltreatment effects were differentiated. Community violence and maltreatment contributed uniquely to conduct disorders, whereas domestic violence did not. Community violence, but not maltreatment and domestic violence, predicted externalizing behavior (McCabe, Lucchini, Hough, Yeh, & Hazen, 2005). However, the sample was not systematically differentiated or stratified by community of residence, limiting implications to how individual variation in exposure affects

these relations and differ by forms of violence.

What Are the Mental Health and Behavioral Effects of Community Violence? The majority of research on community violence has focused on documenting a relation to psychopathology symptoms. Primary interest has been about the effects on externalizing symptoms (aggression, antisocial behavior) and disorders (conduct disorder), internalizing symptoms and disorders (primarily depression but more recently an increased focus on anxiety) and posttraumatic stress disorder (PTSD). Studies have increased in sophistication over the past two decades from convenience samples with cross-sectional correlations to consideration of longitudinal tracking and modeling of how variations in exposure, psychopathology measurement, and covariates and moderators interact in affecting the patterns. Although there are studies that have controlled for exposure to other forms of violence, looked at other risks for psychopathology, and applied careful measurement of type and extent of exposure, these tend to be the exception. For example, a school-based study examining school, home, and neighborhood violence exposure reported that school and home exposure predicted internalizing symptoms, whereas only home exposure predicted delinquency and overt aggression (Mrug, Loosier, & Windle, 2008). As in the other areas of interest, the considerable study of violence exposure due to maltreatment also seems informative about likely effects of community violence (Paolucci, Genuis, & Violato, 2001). In addition, there has been considerable attention to effects of witnessing and being involved in parental domestic violence (Evans, Davies, & DiLillo, 2008; Kitzmann, Gaylord, Holt, & Kenny, 2003). For example, in a sample of youth identified by child protective services as neglected/abused, harsh physical discipline was associated with externalizing whereas witnessing home violence was associated with internalizing behaviors (Maikovich, Jaffee, Odgers, & Gallop, 2008), suggesting that different types of violence exposure may have different effects. These studies could be informative about different levels of exposure. Studies also vary considerably in how recent the exposure was, how carefully chronology of exposure is considered (lifetime versus recent exposure), and correspondingly how time is considered in measuring psychopathology effects (e.g., current status, recent functioning, lifetime occurrence). Also, most studies focus on relative differences in level of psychopathology symptoms, not on presence or absence of disorder. In many cases, the relative differences, although statistically significant, are modest in size between two nonclinical range scores. Yet results from studies that have relied on diagnostic criteria or cut scores thought to correspond to clinical problems are generally consistent with those utilizing symptom scale score comparisons. Similarly, few studies have considered the impact of community violence exposure on practical functioning indicators, such as school functioning, relationship quality, or economic well-being. The few that are available suggest some substantial effects. Most of the existing studies have at least one of the limitations mentioned. These

methodological considerations mean that interpretations are best focused on relative risk increase and probability of disorder than on suggesting direct causal role in mental illness. Similarly, the overall trends across studies provide a more reliable guide than any given study. Methodic reviews have been provided over the past 10 years, including a recent thorough meta-analysis of studies of community violence exposure relation to psychopathology (e.g., Fowler et al., 2009; Lynch, 2003). In addition, a substantial literature has been produced about potential protective factors (Gorman-Smith, Henry, & Tolan, 2004) and about related topics, such as children witnessing domestic and other violence (Evans et al., 2008; Kitzmann, Gaylord, Holt, & Kenny, 2003). These related studies can also inform about the psychopathology effects that seem attributable to community violence exposure.

Meta-Analysis of Effects of Exposure to Community Violence In 2009, Fowler and colleagues reported a thorough meta-analysis of studies of the relation of community violence exposure to psychopathology symptoms. Of the 110 studies and 116 samples (n = 39,667) published before 2005 that had adequate documentation to permit effect size calculation, Fowler and colleagues found that total level of exposure (however measured) predicted mental health symptoms significantly and substantially. The reported effect size averaged d = .63 for externalizing symptoms and d = .45 for internalizing symptoms. In addition, because PTSD symptoms were commonly considered as a distinct outcome in many studies, Fowler et al. examined effects on PTSD symptoms and reported an average effect of d = .79. Effects for Different Levels of Exposure Fowler et al. (2009) were able to examine effects for different levels of exposure. It had often been speculated, and some studies suggested, that those who were direct victims of violence were more at risk for psychopathology effects than those who merely witnessed and/or had heard about violence near where they resided (Martinez & Richters, 1993; Nader, Pynoos, Fairbanks, & Frederick, 1990). For example, Nader et al. (1990) reported that children who were on a school playground during a shooting at that school had more severe symptoms than children inside the building, and both showed more symptoms than children who were not at school that day. Fitzpatrick and Boldizar (1993) found victimization but not witnessing related to depressive symptoms. For externalizing symptoms, direct victimization (d = .78) and witnessing (d = .72) effects did not differ from each other, but both were greater than hearing about community violence (d = .42). For internalizing symptoms, the effects differed by direct victimization (d = .45) compared to witnessing (d = .32) or hearing about community violence (d = .27) but the latter two exposure types did not significantly differ. PTSD symptom effect sizes did not differ by level of exposure (d = .65 for direct victimization, .60 for witnessing, and .63 or hearing about). Effects for At-Risk versus Other Samples Fowler et al. (2009) reported that effects of exposure were not significantly different for externalizing or internalizing problems for samples identified as at-risk by virtue of residence

in high crime communities in comparison to effect for those not having this risk status. However, the relation was significant for both groups on each outcome. When PTSD was the outcome, stronger effects were found for the samples not residing in high risk neighborhoods. Notably, only samples in which the neighborhood risk could be identified were so rated, meaning that some of those samples not so rated simply lacked enough information about where the sample was from to determine if crime was elevated in those communities. Age Patterns in Effects Fowler et al. (2009) reported that some age patterns were significant, comparing adolescentonly to child-only to mixed child and adolescent samples. For externalizing symptoms, correlations were greater for adolescent samples (d = .99) than for mixed child and adolescent samples (d = .53), which were greater than for child-only samples (d = .35). For internalizing, mixed-age samples had the largest effect (d = .51), which was greater than for adolescent samples (d = .45), and which was greater than found for child-only samples (d = .33). For PTSD comparisons, a similar pattern to internalizing symptoms was found, with mixed samples having stronger effects (d = 1.02) than for adolescent-only (d = .66) or child-only (d = .63), but with the latter two not significantly different. Gender and Ethnicity Variations in Effects Gender comparisons in the meta-analysis of Fowler et al. (2009) did not show significant differences in effect sizes for externalizing or PTSD but did find stronger effects for females than males for internalizing symptoms. Ethnicity comparisons showed significant effects for all ethnic groups, with some suggestion of lower effects for internalizing and PTSD for African American and larger effects for PTSD for Latinos when compared with each other and nonHispanic White samples. However, Fowler et al. noted that relatively few studies had femaleonly samples or effects reported in a way that could be separated; likewise, there were not enough studies of Latinos or non-Hispanic Whites to permit sensitive comparisons. Thus, any conclusions need to be drawn with caution.

Example Studies Linking Exposure to Psychopathology Studies Relating Externalizing Problems to Community Violence Exposure Within the overall significant relation between community violence exposure and externalizing symptoms reported by Fowler et al. (2009), there was significant variation in quality of studies and in strength of findings. The interest in violence exposure promoting aggression, conflict with authority, involvement in delinquency, and exhibition of violence has been long-standing (Gorman-Smith & Tolan, 1998; Lambert, Ialongo, Boyd, & Cooley, 2005; Lynch & Cicchetti, 1998). Many of these studies focused on variations in symptomology among those residing in high-risk communities. For example, Gorman-Smith and Tolan (1998) demonstrated that among young men residing in high-poverty, high-crime Chicago neighborhoods, the length of community violence exposure related to level of aggression, even if exposure to family violence and exposure to other life stress were controlled for. As with many other studies,

family functioning (parenting practices, quality of family relationships) explained likelihood of exposure and effects of exposure. Other studies focused on how neighborhood violence level might help explain externalizing symptoms, with residing in such communities considered as representing more exposure. For example, Plybon and Kliewer (2001) differentiated urban neighborhoods using cluster analyses of indicators or relative poverty and crime levels. Children (ages 8–12) living in high-crime, high-poverty neighborhoods exhibited more externalizing behaviors than those living in low-crime, low-poverty neighborhoods. Neighborhood type was mediated by family stress level; and family cohesion moderated the relation of neighborhood type to externalizing. Other studies have examined the relation of community violence exposure to personal behavior. For example, Lynch and Cicchetti (1998) reported that externalizing behavior among urban 7- to 12-year-olds predicted increased exposure to community violence one year later, even when prior exposure was controlled. Gorman-Smith and Tolan (1998) found similar effects but also showed that when controlling for prior behavior, community violence exposure predicted increased aggression. Lynch and Cicchetti also demonstrated that family violence increased likelihood of community violence exposure. Gorman-Smith, et al. (2002) reported similar results for their inner-city male sample. However, in both studies and as examined in some detail by Gorman-Smith et al., the overlap of patterns suggested much variation in likelihood of exposure to violence in the home and also in the community. Not only do these examples show consistent linkage of violence exposure to occurrence of and growth in externalizing symptom levels; they also suggest several aspects of understanding that relation. First, both child behavior and family circumstances contribute to differentiating the likelihood of exposure and effects of exposure. Second neighborhood social relation are important for understanding effects (Sampson, Raudenbush, & Earls, 1997). However, there is still a gap in studies that examine how neighborhood conditions might predict exposure and how such exposure translates to increased aggression and antisocial behavior. Studies Relating Internalizing Problems to Community Violence Exposure Multiple studies have shown that violence effects are not specific to externalizing but usually also affect internalizing problems (e.g., Gorman-Smith & Tolan, 1998; Schwab-Stone et al., 1999). In many cases, the internalizing symptoms were measured broadly, not differentiating anxiety and depression or other internalizing effects. Others focused on one specific type of internalizing disorder. For example, Fitzpatrick & Boldizar (1993) linked depressive symptomology to violence in the community. Lynch and Cicchetti (1998) found that direct victimization (but not witnessing) of violence related to depression symptoms. That study controlled for victimization from child abuse. These findings are in line with the overall finding in the Fowler et al. (2009) meta-analysis showing that effects on internalizing may be more specific to community violence victimization than to overall exposure or merely witnessing violence. Studies Relating Posttraumatic Stress Symptoms to Community Violence Exposure

As noted by Fowler et al. (2009), the largest average effect size of the three psychopathologies considered in the meta-analysis was the relation between exposure and PTSD. As with externalizing, the investigations have gone beyond documenting increased symptomology as a correlate of violence exposure. For example, Garrido, Culhane, Raviv, and Taussig (2010) examined foster care youth to explore the relation between community violence exposure and family violence exposure in predicting trauma symptoms. Community violence was associated with trauma symptoms controlling for family violence exposure, but family violence was not related once community violence was considered. Similarly, in a sample of youth from Cape Town, South Africa, Fincham and colleagues reported that both community violence and child abuse and neglect significantly predicted PTSD symptoms in children. Notably, personal resilience via coping skills moderated the relation of abuse and neglect but not community violence exposure. Gapen et al. (2011) examined how neighborhood processes that are linked to community violence might relate to PTSD symptoms. In their sample of low-income, urban African Americans, perceived neighborhood disorder (a measure of self-reported perceptions of the physical environment) and community cohesion (a measure of perceived social ties) related to self-reported PTSD symptoms even after controlling for previous trauma exposure. Although both neighborhood disorder and community cohesion related to PTSD symptoms, cohesion also partially mediated the relation between disorder and PTSD symptoms.

Studies Relating Other Outcomes to Community Violence Exposure Studies Relating Substance Use to Community Violence Exposure Effects on substance use have also gained some attention, although they are usually secondary to another primary interest, such as effects on PTSD. Significant positive associations have been reported in some but not all studies that tested the relation between exposure to community violence and alcohol and/or other substance use by teenagers (Fagan, Wright & Pinchevsky, 2014; Farrell & Sullivan, 2004; ; Kliewer, 2006; Schwab-Stone et al., 1995; Zinzow, Ruggiero, Hanson, Smith, D., Saunders, & Kilpatrick, 2009). Kaufman (2009) reported a difference in habitual drinking for victimization but not indirect exposure and also no difference for drug use. Similarly, Pinchevsky et al. (2013) assessed direct and indirect exposure to community violence (approximating victimization and witnessing as well as proximity as one combined category) in relation to alcohol and marijuana use. They found each type significantly predicted the use of alcohol and marijuana and binge drinking. However, the effects were weaker and less robust for consideration of covariates for males. In a nationally representative sample, Zinzow et al. (2009) reported that witnessing community violence was associated with substance use, although witnessing parental violence was not. Among those who had witnessed community violence, the chronicity of violence exposure also related to substance use. What may be important, but is often not considered, is that violence exposure is more prevalent in communities in which substance use is less prevalent for adolescents. For adolescents, there is a lower rate of substance use among inner-city adolescents than is found in less-violent suburban communities (Tolan & Morris, 2005). Thus, it may be that among those residing in a given community, there is a positive correlation of substance use and violence exposure, but this effect may occur within a community-level relation of relatively

lower rates of use. Studies Relating Community Violence Exposure to Practical Functioning Effects In addition to the psychopathology symptom effects reviewed by Fowler et al. (2009), it is thought that community violence can affect practical functioning as represented by outcomes such as academic functioning, relationship quality including marital stability, and economic and work functioning (Schwartz & Gorman, 2003). In regard to academic functioning, several studies report lower academic skills and performance related to community violence. For example, Schwarz and Gorman (2003) correlated academic functioning and community violence, noting the relation was partially mediated by depression score. Bowen and Bowen (1999) reported that school danger predicted declines in grades among middle school and high school students and also predicted lower attendance. A specific study of community violence exposure in a large urban sample found that the degree of exposure affected school achievement and engagement (Perkins & Graham-Bermann, 2012; Ratner et al., 2006). Exposure to community violence negatively correlated with IQ and reading ability in a sample of first graders (Ratner, Chiodo, Covington, Sokol, Ager, & Delaney-Black, 2006). ). However, in at least one study of inner-city children, controlling for other stress exposure, community violence did not independently explain school performance (Attar, Guerra, & Tolan, 1994). One longitudinal study noted that poor academic skills predicted earlier exposure to community violence, which related to later level of aggression (Lambert, Bettencourt, Bradshaw, & Ialongo, 2013). Sharkey, Tirado-Strayer, Papachristos, and Raver (2012) reported that children assessed within a week of a homicide that occurred near their home showed lower levels of attention and impulse control and lower preacademic skills than was found otherwise in the community. Parental distress was also elevated, suggesting one mechanism by which this community violence “exposure” could affect child functioning. This smattering of findings suggests there are several ways that such exposure might affect school performance, engagement, and attainment. It might be that violence exposure compromises self-control, attention management, and the executive functioning critical for achievement. It might be that if travel to and from school is perilous, the ability to attend school is decreased. Similarly, it may be that motivation to achieve is lowered when life expectancy is shorter and life is threatened and uncertain. Further basis for expecting such effects can be found from studies of learning processes related to violence (not specifically community violence). De Bellis, Hooper, Spratt, and Woolley (2009) reported lower vocabulary, language processing, and attention control scores related to such violence exposure. Macmillan and Hagan (2004) reported lower academic attainment among adolescent violence victims. Others have shown the same relation for abused children (e.g., Currie & Widom, 2010; Herrenkohl, Sousa, Tajima, Herrenkohl, & Moylan, 2008; Lansford et al., 2005). Not surprisingly, given the consistent relation between community violence exposure and

increased aggression, there is a positive association between community violence exposure and involvement in the criminal justice system. For example, McGee and Baker (2003) reported that victimization by community violence correlated to juvenile offenses in a sample of 12- to 18-year-old African Americans, with a stronger relation for males than females. Among youth already in the juvenile justice system, those previously exposed to community violence were four times more likely to have committed serious criminal behaviors (Preski & Shelton, 2001). In a longitudinal study, Eitle and Tuner (2002) controlled for prior antisocial behavior, other stress, and demographic characteristics and found that exposure to violence as an adolescent predicted young adult crime. Notably, witnessing and direct victimization each independently added to the explanatory capability of this relation. Violence exposure on stability and quality of later relationships has not been studied much. Most of the attention has been on documenting greater violence in personal relationships (Gorman-Smith et al., 2002). Covey, Menard, and Franzese (2013) reported that witnessing community violence correlated to marital status quite modestly (r = .12 to not having married). The relation was at a level comparable to experiencing abuse (r = .14). There has been documentation that adults with histories of child abuse have greater rates of divorce and less stable premarital relationships (Currie & Widom, 2010). One area that has not received much attention is the long-term effect of community violence exposure on employment and income. Prior studies found such relations for exposure to childhood abuse and neglect (Currie & Widom, 2010; Zielinski, 2009) and adolescent victimization (Macmillan & Hagan, 2004; ). For example, adolescent violence victimization was associated with adulthood wages averaging over $1 per hour less than those not victimized. Covey, Menard, and Franzese (2013) tracked the relation between adolescents' exposure to physical abuse, witnessing parental violence, and neighborhood violence in a nationally representative sample of adolescents. Controlling for parental education and SES level, Covey et al. reported that of the three forms of exposure, only community violence exposure related to employment (r = –.08) and income level.

Differentiating or Specifying Effects of Community Violence Exposure As noted throughout this chapter, much of what has been studied about violence exposure has not been focused on community violence or has not differentiated exposure to community violence from other forms. Lynch and Cicchetti (1998) and Tolan et al. (2006) reported that community violence and other forms of violence are associated, meaning those exposed to one form are more likely to be exposed to another. However, both reports emphasize it is a transactional relation, not merely coincidence that is of interest. This means that among those residing where community violence is more prevalent, those with family violence are more likely to be exposed and susceptible to the effects of community violence. It is also likely that there is some increased acceptance and use of family violence when violence is more frequent or normative in the community (Tolan et al., 2006). For example, Cichetti and Toth (1995) compared maltreated and nonmaltreated children in a

longitudinal study in which exposure to community violence was also considered. They found that maltreated children were more affected by community violence, with specific impact depending on type of maltreatment. Others have compared violence effects from home and school to community exposure. For example, Morales and Guerra (2006) examined the relation of internalizing and externalizing symptoms to exposure in community, school, and at home. After controlling for several risk factors for exposure, community exposure was differentiated from home and school exposure in not relating to internalizing or externalizing symptoms. Community violence attenuated the relation between home violence and internalizing and between school violence and externalizing problems. Similarly, there is evidence that exposure across different contexts is not independent, with those more likely to be exposed in one context likely to be affected in another (Laub & Lauritsen, 1998). These studies point out the value of differentiation of violence exposure from other potential sources in identifying potential risk and protective factors. They also highlight the importance of understanding how risk and protective factors may differentially relate to risk for exposure than to exposure impact as well as by extent or type of exposure. Most fundamentally, these findings point to the need to consider interactive dependencies of influence and to trace the transactional patterns that are the basis for community violence exposure effects.

Risk and Protective Factors Associated with Community Violence Exposure and Effects Multiple risk and protective factors have been related to the likelihood of exposure to violence and to the variation in effects of such exposure (Gorman-Smith & Tolan, 1998). These factors range from within-individual characteristics, such as impulsivity and prior aggression level (Lynch & Cicchetti, 2002); to parenting and family relationship characteristics, such as parental monitoring of child behavior and family support and cohesion in managing life challenges (Sheidow, Gorman-Smith, Tolan, & Henry, 2001); to peer influences, such as gang involvement (Thornberry, 1998); to elevated life stress (Attar et al., 1994); to community characteristics related to social functioning of neighborhoods and resources for child development (Tolan et al. 2003). As in other areas of research on this topic, much of what is understood is drawn from inferences from studies of other forms of violence exposure or of community violence considered without controlling for or distinguishing from other forms of violence exposure (Tolan, 2001a). However, there have been empirical investigations of community violence exposure related to indicators within each level of systems of influence in the developmental ecology and also study of the relation of impact if exposed to indicators across several levels. The latter studies often seek to identify protective factors that might mitigate effects or to understand how these additional risk factors might differentially affect those exposed to community violence. Most studies of community violence exposure draw samples from communities that have elevated crime levels. Often these communities are also impoverished because exposure likelihood is also elevated there. In addition, most of the studies available consider how individuals within such neighborhoods differ in likelihood of exposure or in effects from

exposure rather than in how variations in neighborhood conditions systematically relate to group risk. Thus, results inform about individual differences among those residing in communities or attending schools where violence exposure is elevated rather than in the vast majority of communities. Also, as is the case for other topics within community violence effects, often exposure to other forms of violence is controlled for or collapsed into a single indicator. Thus, there is limited information about how community characteristics affect populations residing in communities where violence is elevated, about how community violence specifically might affect those youth, and about how community violence may affect youth residing in communities without elevated violence rates.

Individual Characteristics Linked to Community Violence Exposure Studies of individual differences have concentrated on documenting effects of violence and identifying what differentiates impact among exposed individuals. The most documented individual risk factor is prior exposure (often measured as direct victimization) due to other forms of violence and/or community violence. For example, in a national survey, Finkelhor et al. (2009) reported that children who experienced one type of victimization in the past year had double or even triple the risk of other types of victimization. These risks held true for lifetime exposure as well. Empirical reports of individual differences suggest greater risk for exposure as well as greater impact from exposure for youth with lower self-regulatory capabilities (Lambert et al., 2010; Salzinger, et al., 2002), social-cognitive processes biased toward aggression acceptance and use (Bradshaw, Rodgers, Ghandour, & Garbarino, 2009; Guerra et al., 2003), and higher prior aggression level (Gorman-Smith & Tolan, 1998; Lambert et al., 2005). Hill and Madhere (1996) tested for violence exposure related to social-cognitive features; they determined that perceived level of violence, not outwardly identifiable exposure, that related to socialcognitive features and that this perception mediated the impact of exposure. Overstreet and Braun (2000) demonstrated that perceived exposure may mediate impact of actual level of exposure, suggesting that social-cognitive features may affect not only exposure but also the impact such exposure has on adjustment. Schwartz and Proctor (2000) and Guerra et al. (2003) each reported effects of violence exposure on social cognition (beliefs about aggression use, norms for violence, problem solving when facing conflicts) and that the effects mediated impact on peer relations and subsequent aggression in the respective studies. Notably, in both cases, the correlations were small but significant, despite the studies taking place over some significant length of time. However, neither study tested how these characteristics might affect risk for exposure. Accumulating evidence suggests that exposure to community violence compromises central organizing behaviors (i.e., self-regulatory capacities), such as emotion regulation, and impulse control (Attar et al. 1994; Overstreet & Braun, 2000; Schwartz & Proctor, 2000). Impairments in self-control or behavioral control may limit access to social support or increase likelihood of victimization (Pope & Bierman, 1999; Schwartz & Proctor, 2000). Low and Espelage (2014) found that impulsivity related to more exposure and exacerbated exposure effects by

increasing involvement with antisocial peers. In addition, there are indications that prior psychopathology symptom levels relate to greater likelihood of exposure. Gorman-Smith and Tolan (1998) reported that prior depression and prior aggression related to subsequent violence exposure among a sample of inner-city highrisk males. Kliewer (2006) and Overstreet and Braun (2000) reported a similar relation to trauma symptomology. All of these studies noted that psychopathology did elevate risk for exposure but also that greater exposure related to a relative increase in symptoms as well. Thus, similar to the transactional model suggested by Lynch and Cicchetti (1998) for the relation between child maltreatment and community violence exposure, it appears that aggression and depression relate both to greater exposure and also to greater effect if exposed. Similar but more complex findings were reported by Boyd, Cooley, Lambert, and Ialongo (2003). In that study of urban students from economically disadvantaged schools, teacher and parent reports of aggression at baseline each predicted subsequent exposure to violence, but only for boys. In addition, impact on later behavior was dependent on the boys' anxiety level, with an effect only for low-anxiety boys. In longitudinal follow-up with this urban student sample, Lambert, Ialongo, Boyd, and Cooley (2005) reported that among males, aggression, anxiety level, and depression each related to subsequent victimization but not exposure due to witnessing. However, for females, the relations held across type of exposure. Overall, the reports suggest that effects of exposure, including variation in exposure type effects, immediacy of effects, and psychopathology symptoms affected, may be different by gender. Studies of the impact of moderators of violence exposure have identified coping skills as a potential mitigating factor (Tolan et al., 1997). Dempsey (2002) showed that violence exposure relates to reliance on aggression and other negative forms of coping. Edlynn, Gaylord-Harden, Richards, and Miller (2008), found methods used for effectively coping with violence exposure may be considered maladaptive in other contexts (e.g., the use of avoidance). In a study of anxiety impact of violence exposure, Edlynn et al. found that direct problem solving had no moderating effect for young men residing in high-violence, highpoverty urban communities. Other studies suggest that the benefits of coping may be dependent on adequate social support, particularly from adults (Hammack, Richards, Luo, Edlynn, & Roy, 2004) ). Tolan and Grant (2009) summarized the research on coping for youth residing in inner-city communities—the communities where violence exposure is most elevated—as often requiring strategic use of methods such as avoidance and reliance on others for protection and guidance largely because personal control over one's own circumstance may be limited. Direct action and individual problem solving may lead to decreased safety and/or engulfment in social challenges that are developmentally beyond what is manageable by children and adolescents. For example, if faced with an unfolding violent situation, trying to stop it may increase the likelihood of victimization. At the same time, reputation and subsequent safety might rest on not opposing the violence. Others have suggested that coping skills for violence exposure need to be understood from a critical consciousness framework in which the political and social forces thought to be the reason for elevated violence exposure in a given community are discussed and utilized to help

support productive coping (Bentley, Thomas, & Stevenson, 2013; Watts, Abdul-Adil, & Pratt, 2002). From this framework, coping is based in cultural identity, awareness of inequities, and positive alternatives to frustration and aggression or depression.

Parenting and Family Characteristics Exposure to community violence has been related to family structure (Martinez & Richters, 1993), although inconsistently (Fitzpatrick & Boldizar, 1993). Miller et al. (1999) noted that younger siblings of delinquents were more likely to be exposed to community violence. When differentiated from family violence exposure, there is no relation of family structure (e.g., oneparent family) and violence exposure (Hanson et al., 2006). Some studies have shown a relation between parenting skills and violence exposure as well as impact of exposure (Low & Espelage, 2013). However, the majority of studies that have differentiated particular parenting practices or qualities of family relationships find no significant relation to exposure and show inconsistent relation to impact of exposure (GormanSmith & Tolan, 1998; Miller et al., 1999), except as related to maltreatment (Lynch & Cicchetti, 1998). Risk for exposure to community violence has been related to family alcohol or drug problems (Hanson et al., 2006; Stevens et al., 2005). Similarly, children from families where there was intimate partner violence showed increased risk for community violence exposure (Lynch & Cicchetti, 1998; Rumm et al., 2000). These findings suggest that a breakdown in parenting may result in greater exposure risk for children, and they may incur additional accompanying risk for effects due to similar lack of access to family as a resource (Tolan et al., 2013; Tolan, Sherrod, Gorman-Smith, & Henry, 2004;). Some have suggested that apart from the effects attributable to parental and family violence exposure, family influence on risk for exposure may be better understood as violent events causing a disruption in parenting support and resources (Tolan & Gorman-Smith, 2002; Trentacosta et al., 2009). From this perspective, violence in the community creates challenges that strain family capabilities, imposes threats that overwhelm or exacerbate other stress faced by families in these communities, and/or occurs along with many other risk factors such as violence-involved peers that also challenge parental efficacy (Overstreet & Mazza, 2003; Salzinger, Feldman, Stockhammer, & Hood, 2002; Tolan & Gorman-Smith, 1997). A study by Gorman-et al., (2004) showed an exception to limited indication of parenting and family relationship correlates to exposure. That study examined person-based patterns of combined parenting practices (monitoring, discipline consistency, use of positive parenting) and family relationship characteristics (cohesion, support, organization) to classify families into four groups: exceptionally well-functioning (high on all scales); typical families (midrange on all scales); struggling families (low on all scales); parenting-focused families (mid to high on parenting scales but not on family relationship scales). Struggling families were more likely than the other types to have sons with greater exposure to violence, but both struggling and parenting-focused families showed greater impact of such exposure than the other two types. The results suggest that children from typical and exceptional families were protected from the

effects of exposure on subsequent aggression and depression. This difference in findings for the person-oriented configuration of characteristics from that obtained from variable-based correlations seems to reflect how patterns of characteristics might represent family risk that are not captured in variable-based studies. Specifically, these results suggest that overall functioning of the family and the cumulative impact of different configurations of family relations are important, not just specific, characteristics. In addition, this study suggests subtle differences in what might relate to exposure and what might relate to effects. Spano, Rivera, and Bolland (2010) recently reported a similar approach and found that particularly vigilant monitoring was related to less exposure of urban impoverished youth but that it had a less clear protective effect for impact. These results are consistent with the studies of Gorman-Smith et al. (2004) and others of family relationship qualities as protective of violence exposure effects (Hammack et al., 2004; Hardaway et al., 2012). In particular, the prevailing findings seem to be that family functioning that includes strong monitoring and warm relationships between parent and child can moderate effects of violence exposure even if it is unable to prevent such exposure (Kliewer, Murrelle, Mejia, Torres, & Angold, 2001).

Peer Influences Multiple studies have documented a positive relation between involvement with delinquent or violent peers and victimization from community violence (Fagan, Piper, & Cheng, 1987; Felson, 1997; Lauritsen & Davis-Quinet, 1995; Tolan et al., 2003). Affiliation with delinquent peers is related to likelihood of witnessing community violence (Halliday-Boykins & Graham, 2001). Children who associated with more deviant or delinquent peers in one year have greater exposure to community violence in the next year (Lambert et al., 2005; Salzinger et al., 2002). Such association seems to promote risk particularly for high-risk youth. For example, Lambert et al. (2005) reported that deviant peer affiliation moderated the association between aggressive behavior and exposure to community violence. It would be important to disentangle how community violence exposure is related to engagement with delinquent or violent peers (Thornberry, 1998). One important distinction is to differentiate how antisocial tendencies may lead to affiliation with peers who are more violent, thus increasing community violence exposure, from how such exposure may spur peer affiliations with deviant youth. For example, it is evident that violence perpetration increases as youth get involved in gangs and decreases if they leave gangs (Thornberry, 1998). The theoretical explanation is that by being affiliated with violent youth gangs, use of violence is promoted and the youth tend to exhibit violence they might otherwise not perpetrate to. It seems likely that community violence exposure could have similar processes of effects, but it is also plausible that such violence has less direct and immediate influence than that related to peer relationships. More recently, studies of protective factors for violence exposure effects have identified social support and in particular engagement with peers for social support. For example, Hammack et al. (2004) found that social support (mostly from peers) emerged as a protective-stabilizing force for witnesses of violence both cross-sectionally and longitudinally. Similarly, Hardaway, McLoyd, and Wood (2012) reported that engagement in prosocial activities moderated violence exposure, an effect they thought might be related to support from prosocial peers.

Notably, these findings indicate that support, not only from peers but also from family and the family-peer network, serves to mitigate exposure impact (Kaynak, Lepore, & Kliewer, 2011).

Neighborhood Social Relations and Structural Characteristics Community violence exposure is defined in part by proximity to where the violence occurs. Also, it is well documented that violence exposure is greatest in the lower-SES communities in large cities. Therefore, it seems likely that much of the risk for exposure, and potential solutions to reduce exposure, rest in understanding neighborhood or community characteristics like these (Tolan, Sherrod, Gorman-Smith, & Henry, 2004). ). What is now emerging is a theoretical framework for understanding neighborhood characteristics and processes affecting development that can be applied to understanding violence effects (Henry, Gorman-Smith, Schoeny, & Tolan, 2014; Sampson et al., 1997; Tolan et al., 2004). This framework focuses on how community structural characteristics, such as economic resources, income level of residents, and single-parent family prevalence, can affect social problems, such as high crime rates and limited opportunities for conventional achievement. These issues are in turn related to social processes of neighborhoods, such as collective efficacy, informal social control, shared norms, and felt responsibility for the neighborhood safety level. The interest in how neighborhood differences in economic and social resources and problems relates to crime and particularly to crime inducement has been long-standing (J. Wilson, 1987). For example, community-level variations in population makeup, home ownership, mobility, and housing density positively relate to variations in community-level crime (Bursik, 1988; Byrne & Sampson, 1986). In these and subsequent studies, structural characteristics were differentiated from social characteristics or processes (Sampson et al., 1997; Tolan et al., 2004; J. Wilson, 1987). The term structural characteristic refers to demographic, economic, and political differences that mark communities economically and politically and are related to resource limitations and problematic outcomes (Tolan et al., 2004). Structural characteristics tend to be examined in terms of the average economic status of residents and the range of income levels within the community, the crime rate, and the opportunity for employment and adequate housing and education. Other structural characteristics, such as quality of schools and social and health resources, can also be related to violence (Henry et al., 2014). Theoretically, structural limitations affect the opportunity for neighborhood social processes that are thought to be important for healthy development. The term neighborhood social processes refers to relations among residents (Sampson et al., 1997; Tolan et al., 2004). These include processes such as informal social control or how much neighbors regulate each other's behavior through expectations, behavioral norms, and interaction. The classic example of informal social control is the extent to which youth in a community might be subject to control of deviant behavior by the adults in the community (Sampson et al., 1997; Tolan et al., 2004). These processes have been incorporated into an overall construct referred to as collective efficacy (Sampson et al. 1997), which refers to a felt, shared ability among residents to manage issues and problems facing the community. Multiple studies have tracked the combined and interdependent effect of these different aspects of community ecology to violence perpetration, including increased exposure to violent peers in the community (Chung & Steinberg, 2006;

Tolan et al., 2003). Recently, Henry et al. (2014) validated a set of scales that measured three neighborhood processes commonly referred to in the literature and that relate to child development outcomes: norms (type of behavioral expectations and extent shared), informal social control (regulation of neighbors by group expectations and felt responsibility for community safety and order), and social connection (extent to which residents organize to manage the neighborhood, spend time together, feel connected). Notably, these sets of scales were validated using neighborhood-level scale construction and also focusing on child development as well as crime outcomes. Many prior measures actually were constructed and validated as individual level measures, even though they were used and interpreted as representing community-level constructs. Although violence exposure seems to be concentrated in high-poverty urban neighborhoods, the extent of violence and exposure of youth can vary substantially among those with similar structural characteristics (Tolan et al., 2004). Browning, Gardner, Maimon, and Brooks-Gunn (2014) reported that exposure to community violence that was life-threatening had effects on psychopathology for those youth in neighborhoods with low collective efficacy but not when efficacy was elevated. Tolan et al. (2004) showed that while neighborhood processes could be predicted by structural characteristics, structural characteristics also mediated the relation of neighborhood processes to exposure of violent peers and subsequent risk for violence perpetration. Sheidow et al. (2001) found that neighborhood social process level had a small effect on violence exposure among young men residing in poor urban communities, but youth from families with a combination of lower parenting skills and less supportive family relationships had particular risk. The most direct test of neighborhood processes related to youth community violence exposure was a study by Fagan, Wright, and Pinchevsky (2014) that showed that collective efficacy moderated the relation between community violence exposure and substance use by adolescents. This finding suggests that neighborhood processes can serve a protective role against the structural characteristic of high levels of violent crime. However, the limited number of studies points to how little we know about what neighborhood characteristics, beyond poverty, relate to risk and how social processes might help mitigate exposure and effects from exposure. These initial studies suggest that a critical area for empirical investigation is how crime rates that may drive exposure to community violence are related to neighborhood social characteristics within those communities with structural characteristics that also relate to crime rates.

Theories about How Community Violence Exposure Causes or Increases Individual Risk for Psychopathology Varied explanations have been offered regarding how community violence exposure is harmful. Most explanations have not differentiated community violence exposure from violence exposure in general, or they are drawn directly from formulations of how family violence might affect a child (Margolin & Gordis, 2000). However, an additional thread of reasoning compares residing in communities with high violence rates to exposure to war conditions,

natural disasters, or horrific accidents and maltreatment (e.g., kidnappings) (Fitzpatrick & Boldizar, 1993). These explanations tend to emphasize a trauma process of impact that either precipitates pathology or exacerbates prior risk into expression of syndromes. Other theories have focused on attachment disruption, similar to what is thought to be an effect of neglect and abuse (Cichetti & Toth, 2005). Social-cognitive theories have been offered that suggest the impact is to shift normative expectations about behavior and of rewards or functionality of violence. Still others have suggested that violence exposure is often part of an overall ecology of exceptional stress and unpredictability and so is best understood as one of many challenges faced by children and families residing in impoverished urban communities. In this approach, violence exposure is considered as distinct from, but one of many, stress-inducing experiences affecting risk (Tolan & Gorman-Smith, 1996). Explanations framed using this perspective tend to see the effects as cumulative and linear, with more exposure and less stress-coping resources equating to more symptomology.

Explanations Emphasizing Trauma Process Trauma-focused explanations rest on the concept that severe fright and disgust caused by witnessing violence, being directly victimized, or residing in an unsafe community overwhelms and perverts neurophysiological functioning—particularly attention regulation and inhibition of impulses. Further, this traumatic exposure also impedes or misdirects further development of these key systems that increases susceptibility to other psychopathology risk factors or sets overactive or underactive attention and arousal management and the relation between emotion and behavior. Overdeveloped fear conditioning is seen as a key process, with additional effects on hypothalamic-pituitary-adrenocortical (HPA) axis regulatory systems. Both processes are thought to either lead to hypervigilance about threat (and related anxiety, depression, or aggression) or problematic habituating and underreactivity to violence (and related aggression and depression) (Horowitz, McKay, & Marshall, 2005; Perry, 2008). Children who are still developing are particularly susceptible to these effects and also suffer from the ongoing unpredictability of the environment, including a lack of ability to rest secure and safe (Garbarino, Dubrow, Kostelny, & Pardo, 1992). At too early an age and in an ongoing, clearly unhealthy way, these children must be always on alert. As a result, psychological functioning and underlying neurobehavioral regulatory systems are compromised, analogous to the impact of war exposure found in veterans and survivors of disasters. This may lead to anxiety, depression, as well as use of aggression as a means of intended self-protection (Tolan et al., 1997). A related theory is that violence exposure in the community disrupts attachment because it brings into acute awareness the inability of caregivers to provide safety and belonging, which are so critical to healthy development. Most essentially, the precariousness of a safe environment and dependable care is thought to disrupt secure attachment capability, which acts as a trauma to undercut relationship engagement, trust, and communication with parents and other caregivers and the ability to properly depend on others. All of this exacerbates risk for psychopathology and/or is thought to cause it (Lieberman, Chu, Van Horn, & Harris, 2011; Perry, 2008). This approach is different from the framework that says experience overwhelms

neurobehavioral regulatory systems. Instead, it posits that the mechanism of psychopathology is disruption of social relationships through effects on personal attachment and relationship skills. Although these trauma frameworks provide important ideas and potentially critical insights about how community violence affects children, they are primarily based on evidence accumulated about child abuse and neglect or domestic violence. An exception is a study by Farrell and Bruce (1997) that tracked community violence exposure to emotional distress features related to trauma. That study focused on individual differences in exposure and found a modest positive link. However, the authors did not focus on documenting a traumatic process as the cause of behavior change, but rather to correlate how extent of symptoms thought to represent post traumatic stress disorder corresponded to violence exposure extent. Thus, it seems likely that the same processes implicated for other forms of violence exposure might be at work when people are exposed to community violence. However, as with many other theories, there has been almost no tracking of these effects specific to community violence exposure. In part this is because the needed studies would entail comparing samples defined as community violence exposed to counterparts who have only had traumatic events or family violence experiences or to nonexposed children who are otherwise similar (e.g., also reside in low-SES urban communities). Creating such a study and garnering the resources to enable measuring the neurobiological effects and related symptomology presents numerous challenges. This lack of empirical studies has limited how much can be inferred from findings of violence effects from abuse, witnessing domestic violence, or disaster experiences (Cicchetti & Toth, 2005).

Social-Cognitive Impact Theories of Community Violence Exposure Effects Social-cognitive theory has also been offered as an explanation of community violence exposure but almost exclusively in relation to subsequent aggression levels. Within this framework, violence exposure promotes aggressive reactions but also supports the use of aggression as a method of managing fear, frustration, and conflict (Dodge & Somberg, 1987). Growing up in a community where violence exposure is more likely means that it is more likely to be seen as normative, acceptable, and useful (Guerra, Huesmann, & Spindler, 2003). For example, Guerra et al. (2003) reported that community violence exposure had a significant effect in increasing aggression, normative beliefs about aggression, and aggressive fantasy. In older children it showed effects on social-cognitive processes. Additionally, this study showed that these changes in social cognition mediated the effects of exposure on aggression exhibition. While emphasizing the role of exposure in increasing an expectation of violence from others and to form a bias toward viewing motivations of others as hostile (which increase use of violence), these social-cognitive processes have also been tied to depression symptoms (Dodge, 1993). Social cognitions related to conflict, frustration, and self-control are shaped by imitating others (who are also likely to use and approve of violence) by the utility of violence in enabling development (e.g., creating status, reducing threat, accessing rewards) and shifting what is

considered normative or acceptable behavior. In turn, this misdirected form of self-control, self-image, and consideration of others in forming behavioral responses leads to a more limited arsenal of alternatives to aggression, more use of aggression, lack of impulse control over life span, and less reward or reward orientation related to nonaggressive behavior (Morales & Guerra, 2006). Notably, this approach has been based on empirical studies that suggest correlation among the components and some moderating and mediating relations (Attar et al., 1994; Guerra et al., 2003) ). For example, Chaux, Arboleda, and Rincon (2012) corelated residence in more or less violent communities in Bogota, Columbia to aggression. The 1,235 fifth- to ninth-grade students studied reported a relation between exposure to community violence and levels of proactive and reactive aggression (aggression used to solve a problem versus used in reaction to immediate threat). Moreover, normative beliefs supporting use of aggression, hostile attributions of intent, positive expectations from use of aggression, and low guilt about using aggression each partially mediated that relation. Notably, much of the theorizing about violence effects has been formulated through studies of youth residing in high-risk communities and through measurement of variation in individual behavior based on extent of exposure (Guerra et al., 1995; Guerra, et al., 2003). These theories also correspond to stress models in that exposure to violence is seen as one of many forms of environmental challenge that might affect normative beliefs about behavior, particularly the role of aggression in conflict management (Cooley-Quille et al., 2001; Guerra et al., 2003 ). Thus, some models link social-cognitive processes to a developmental-ecological model that characterizes exposure as inducing stress in parallel, linked, or contributing to a confluence of pathology-promoting process (Morales & Guerra, 2006).

Stress Models of Community Violence Exposure Effects The primary supposition of this approach is that violence is one form of life stress, a particularly powerful and potentially virulent form (Gorman-Smith & Tolan, 1998). Based on the observation that community violence exposure is often based on residence within communities that are marked by many other stressful conditions and families facing these also have limited social and economic resources for managing stress, it is suggested that violence exposure may result in anxiety, depression, PTSD symptoms, and aggression (Tolan et al., 1997). This theorizing has focused at the level of individual coping skills and resources, such as social support, parental guidance, and protection from places and situations that might increase community violence exposure. Linkage to neurodevelopmental processes in stressregulatory systems has been limited, and most of these studies have not considered other forms of stress. An exception is a study by Sharkey et al. (2012), who assessed children within a week of when a homicide occurred near their homes to test for its effects on child behavior, utilizing a stress framework. They found lower levels of attention and impulse control and lower preacademic skills than found otherwise in the community. Parental distress was also elevated, suggesting that one mechanism by which this community violence exposure could affect child functioning is that it renders violence-exposed parents less able to help to provide needed regulation and care. Only a few studies have examined the impact of community violence exposure on physiologic

markers of stress response in children (Kliewer, 2006; Wilson, Kliewer, Teasley, Plybon, & Sica, 2002;). For example, Kliewer (2006) examined the association between salivary cortisol levels and exposure to violence in 101 African American youth. Witnessing community violence was associated to lower baseline cortisol and increased cortisol in response to a mild stressor (e.g., watching a video showing scenes of community violence). Much of the theorizing about violence exposure treats it as a life stress that can overwhelm coping mechanisms and support resources and strain the physiological systems involved in managing stress; in doing so, it induces distortions in psychological functioning that characterize posttraumatic stress (anxiety), depression, and aggressive tendencies. For example, Attar et al. (1994) showed that exposure to violence and threats exacerbated the effects of more minor stress. In a related study, Gorman-Smith and Tolan (1998) documented that for inner-city youth, violence exposure has a distinct effect on level of externalizing and internalizing symptoms, over and above that attributable to general life-stress level. Although it helped explain symptoms from other forms of stress, it was distinct from being just another form of stress. In that study, the effect was found even after accounting for individual variations in behavior that might have prompted different levels of exposure to violence. White, Bruce, Farrell, and Kliewer (1998) reported a similar set of effects, with community violence promoting additional effects over and above those found from other forms of stress. Common to these diverse theoretical explanations is a recognition that exposure to violence is related to larger developmental context characteristics, such as residing in inner-city communities, or transactional confluence processes, such as low parental monitoring. Similarly, residing in such communities may increase opportunity for proximity to violence or affiliating with those likely to perpetrate violence (Lynch & Cicchetti, 1998; Tolan, Guerra, & Kendall, 1995). These developmental-ecological models emphasize that multiple risk processes, with interdependencies among the processes, are implicated in both the risk for exposure and the impact a given level of exposure may have. For example, violence exposure is more likely to occur to families with fewer economic and social resources and in neighborhoods with fewer social connections among neighbors but also to those youth with more impulse problems and more antisocial friends. The limited evidence does not suggest that risk is explained simply by the accumulated adverse experiences in life; rather, it may be due to some particular interactions and may vary substantially by the type of community in which one resides. At the same time, there is growing interest and basis for understanding how violence exposure gets under the skin to affect psychopathology through neurodevelopmental processes.

Neurodevelopmental Processes that Might Be Implicated in Community Violence Exposure Neurodevelopmental explanations of community violence exposure are essential to translate the social and behavioral nature of these experiences to individual developmental psychopathology effects. As with all areas of developmental psychopathology, the relation

includes a dual consideration: neurodevelopmental functioning and processes that differentiate susceptibility to environmental conditions and experiences and determination of how the environmental experience (in this case community violence) affects brain and neural physiology, systems, and functioning. Much work has been conducted to identify the impact of violence on neurodevelopment, but little has focused on community violence specifically (Cicchetti & Toth, 2005; Mead, Beauchaine, & Shannon, 2010). Most understanding of neurodevelopmental effects from violence exposure has come study of abused and neglected children. These samples often included a mix of children who were exposed to violence and those neglected but not exposed to violence. In addition, as exposure to abuse and neglect are also related to probability of community violence exposure and other failures in care and safety, it is difficult to parse what might be effects of abuse/neglect, community violence exposure and attendant failures in care, attachment, other stress, and co-occurring insults to the neurodevelopmental systems and course (Lynch & Cicchetti, 1998; Malik, 2008). Yet it may also be that violence exposure source or type does not significantly affect the neurodevelopmental processes or susceptibility to pathological effects on development. Thus, while providing many potential clues to the likely processes of and effects from community violence, the applicability of neurodevelopmental understanding gained from studies of violence impact is not yet determinable.

Genetic Contributions to Violence Exposure Susceptibility Gene–Variant Contributions from Other Forms of Violence Exposure Genetic contributions might relate to violence exposure in many ways, but these can be grouped into two main ways (Tolan et al., 2013). First, genetic differences related to tendency to be aggressive or to be antisocial may lead to greater likelihood of being involved with community violence (Raine et al., 2005). Second, there may be genetic differences in disposition that are affected differently by violence exposure. Several studies have suggested that genetic polymorphisms (e.g., monoamine oxidase A [MAOA]) moderate the relation between violent abuse as a child and violence and aggressive outcomes (Caspi et al., 2002; Kim-Cohen et al., 2006). Others have traced genetic variations (serotonin transporter; 5-HTT) in response to stress (Cicchetti, Rogosch, & Sturge-Apple, 2007). Therefore, it is likely that there are some genetic variations in response to community violence and in the outcome of such exposure. However, since specific genetic studies of community violence exposure effects have not been conducted, it could also be that the process and the genes involved are not analogous to that of direct victimization through abuse (Huizinga et al., 2006). Other gene variants and genes might be involved. It is likely that not only will there be multiple genetic variations implicated but also that those identified as most likely candidates may not prevail. It seems likely that gene variants will help explain the differential effect of violence exposure and may help explain the differential likelihood of exposure. It is also likely that networks rather than specific gene variants will be implicated in this effect and that a confluence of genetic contributors and other developmental influences will also contribute.

One possibility is that exposure to community violence is not interactive in the sense of only having influence in the presence of certain genetic variants but that there are two main effects or a gene–environment (G×E) correlation. For example, a very severe level of abuse exposure is the environmental moderator implicated in several G×E interaction demonstrations of violence effects on aggression (Kim-Cohen et al., 2006). Below that level, the effects seem to be additive, not interdependent. For example, Foley et al. (2004) utilized a lower threshold of violence exposure in a G×E interaction study of violence effects, including exposure to witnessing domestic violence, which may be more analogous to community violence exposure. They report a main effect for this exposure on conduct disorder in addition to moderation by genetic vulnerability. The relation also could be a passive or evocative correlation. Those with the tendency to be violent or to engage in social relations with other aggressive individuals may be more likely to be exposed to violence. Those with a genetic disposition toward antisocial personality characteristics may be more likely to evoke parental hostility or other risk, thereby exacerbating related environmental responses related to exposure. Since these studies are related to family-based violence (maltreatment), the coalescence of multiple methods of genetic influence are also present, which may not transfer to how community violence exposure and effects occur (Tolan et al., 2013). Similarly, maltreatment and abuse have been related to depression likelihood through polymorphism studies, primarily in the promoter region of 5-HTT, the serotonin transporter gene (Cicchetti et al., 2007). This study also reported that different allele variations had a G×E effect depending on the type and extent of maltreatment exposure (e.g., neglect, sexual, physical). The effect also related to stress level. More generally, studies of depression have focused on life-stress experience, including violence exposure, showing stress exposure moderates a genetic tendency for depression, increasing risk more for those with heritable risk than those without (Kendler, Kuhn, Vittum, Prescott, & Riley, 2005). For example, Kaufman et al. (2004) found that among a sample of maltreated children, the short/short allele variant related to greater probability of current depression than either the short/long or long/long variant. However, they also noted that this relation was moderated by social support, which is often viewed as protective against stress. The trend across these G×E interaction studies of maltreatment and stress and depression suggests a more complex set of interactions than implicated in studies of antisocial behavior outcomes. These findings seem consistent with a stress interpretation of violence impact on developmental psychopathology (Gorman-Smith & Tolan, 1998). The G×E interaction studies suggest two processes for community violence exposure effects. Exposure leads to greater likelihood of aggression for those with the at-risk allele MAOA pattern and to depression and other internalizing problems for those with the risk allele for 5HTT. Susceptibility to type of effect also could have some variability by gender. Some have speculated that MAOA and aggression relation and 5-HTTT and depression relation moderated by violence exposure differs by gender, with males primarily affected by the MAOA and females by the 5-HTTT (Mead et al., 2010). Whether this is borne out in additional studies of violence exposure or is tested and found applicable to community violence exposure remains to be seen. A dual neurodevelopmental process of violence impact

is suggested: increasing aggressive tendencies by breaking down self-regulation systems and increasing depression by overwhelming stress management systems (Crockett, Clark, Hauser, & Robbins, 2010). It may well be that rather than two distinct G×E interactions that affect impact of community violence exposure, the interactions could act together. It may be that, if moderated by genetic susceptibility, the effects of violence exposure act through stress systems and also through threat-impulse control systems. Thus, effects may vary from internalizing and externalizing to the extent that both self-control and self-regulation of emotion are susceptible. What is not determinable is how applicable the results of these studies are for community violence when the evidence for G×E effects are only from exposure to family violence, usually as direct victim, and with substantial levels necessary for interaction. In addition, the evidence of the MAOA polymorphism and family violence exposure has been replicated only for White samples, although many studies have been limited in ethnic diversity. For example, Widom and Brzustowicz (2006) failed to find this moderation for the non-White portion of their sample. Also, 5-HTT polymorphism relating abuse to depression has been inconsistent in findings across genders and age groups. Given that community violence exposure disproportionately occurs for minority persons, this finding suggests that further caution be exercised on generalizing from these findings. Although samples used in these gene expression studies have been more diverse ethnically than those demonstrating MAOA interaction, there has not been the type of comparison that might permit consideration of how ethnic group differences in gene variants might affect impact. Moreover, although there have been demonstrations of poverty relating to cortisol differences, it is not clear how this relates to G×E interaction (Lupien, King, Meany, & McEwen, 2001). Since violence exposure is related to poverty and residence location and since it is also a correlate of ethnic group in our current society, these contextual differences may substantially shift the relation of environment and genetic susceptibility. Given that community violence exposure disproportionately occurs to non-White persons, this finding suggests further caution in any generalizations from these findings. Genetic Variation in Environmental Sensitivity An alternative to a polymorphism variation risk susceptibility model is the formulation that emphasizes sensitivity to environment as a key genetic difference: that some persons have a tendency to be more affected by environmental conditions, good or bad, than others (Belsky & Pluess, 2009). One relevant example noted by Belsky and Pluess (2009) in support of this formulation of G×E effects is that toddlers who scored high in proclivity to get angry at 15 and 28 months were more likely than other children to manifest behavior problems at age 5 if provided ineffective parental guidance and were less likely to display such problems than others if parental guidance was effective. It may be that those most affected by community violence are those most sensitive to environmental disorder, fear-inducing threats, or the shock of experiencing, seeing, or living with violence. Although each of these models might be quite important in understanding how individual differences in reaction to violence exposure occurs, at present, a G×E study of community violence exposure has not been conducted. Thus, little is known about how G×E relations should be considered in psychopathology formulations of community violence exposure.

Epigenetic Explanations of Violence Impact Another genetically informed area of study of violence effects might show the importance for studies of community violence impact. This is the rapidly developing area of telomere science. This field focuses on cellular and molecular processes affecting telomere length, particularly related to stress and stress-related aging processes over the life span. Shorter telomere length is associated with advancing chronological age and also increased disease morbidity and mortality. Thus, interest has been stimulated in how stress might accelerate erosion of telomeres, even early enough in life to affect newborn setting of telomere length. Shalev et al. (2013) tested for relation stress to telomere erosion or aging in a prospective-longitudinal study in children and found that children who experienced multiple forms of violence exposure (either exposure to maternal domestic violence, frequent bullying victimization, or physical maltreatment by an adult) showed significantly more telomere erosion in buccal cells in measurements between ages 5 and 10. This effect remained even with controlling for multiple potential confounds. This finding suggests that perhaps violence exposure acts to age or to deteriorate telomere length, which is thought to explain susceptibility to psychopathology. From the findings of G×E relations and apparent interactions, it can reasonably be concluded that there may be genetic variations in susceptibility to community violence and genetic impact that is analogous to what has been demonstrated for other forms of violence exposure. Moreover, it appears that exposure can have effects through stress regulation and self-control or aggressive tendencies in the face of threat influence. In addition, G×E models (or gene network and multiple environmental contributors) that emphasize specific risk susceptibility may provide understanding. Or these models may affect relative sensitivity to environmental conditions. These studies may provide important models, suggested processes, candidate gene and environmental precipitants, and insight about brain plasticity related to psychopathology effects and therefore guide helpful preventive and treatment interventions (Cicchetti & Curtis, 2006). However, most fundamentally, studies are needed that explicitly and distinctly consider community violence.

Alteration of Neural Networks, Brain Areas, and Regulatory Systems Violence exposure effects on psychopathology and functioning are presumed to occur through some alteration of neural physiology, processes, and regulatory systems that affect symptom occurrence and functioning. Most essentially, it is believed that shocking, threatening, unpredictable, and sensory-overloading experience through violence disrupts healthy developmental processes for adaptation by altering brain systems that manage threat-reaction and stress responses. Exposure to violence, particularly early in development, is seen as potentially altering formation, efficiency, and proper functioning of important areas of the brain. These inferences are drawn from studies of direct victimization, usually from caregivers. However, some studies of exposure to domestic conflict and violence have been influential on this conceptualization. Unlike neglect, which is viewed as failure to consistently stimulate in order to promote

organization and relative stability in emotional functioning, violence exposure is unregulated, unpredictable overstimulation, with harm resulting from repeated and extreme overstimulation. As suggested by G×E interaction studies, this overstimulation can occur through multiple systems and can affect orienting and regulatory functions of the cortex (cognition), limbic system (affect regulation), diencephalon (e.g., fine motor regulation, startle response), and brainstem (heart rate, blood pressure regulation). Since each system is affected by experiencedependent neurochemical signals and each has different periods of sensitivity, the impact can vary as to which system is compromised and how that affects development. Also, since many of these systems can be affected by in utero experience, children may be born with substantial variation in sensitivity to disruptions such as violence exposure (Perry, 2008). For example, in animals, experimental inducement of prenatal and perinatal stress can alter hippocampal organization and the HPA axis and have lifelong effects (Plotsky & Meaney, 1993). Lower Areas of Brain (Brain Stem, Diencephalon, Hypothalamus) Lower areas of the brain are the portals for neural/sensory input and relay sensory information to the primary sensory receptive areas of the cortex into a feedback process of sorting and actions related to threat experience. The reticular activating system (RAS), the hypothalamus, and the thalamus are brain regions most likely to be affected through violence exposure. The RAS is a network of arousal related neural systems through which neurotransmitters, including epinephrine (EPI), norepinephrine (INN), dopamine (DA), and serotonin (5-HT), orient and set regulation of arousal, fear, stress response, and threat reaction. The hypothalamus and thalamus receive input from various afferent sensory systems and transmit arousal information from the RAS to the limbic, other subcortical, and cortical systems (Castro-Alamancos & Connors, 1996). Violence exposure could affect the lower brain areas through affecting orientation sensitivity, with primary focus and evidence to date on startle response. For example, several studies have identified exaggerated startle responses among people with PTSD related to war (Morgan, Grillon, Southwick, Davis, & Charney, 1996) and among female sexual assault survivors (Morgan, Grillon, Lubin, & Southwick, 1997). Not all studies have shown this link, however. In some cases, the startle response is blunted. The implication is that exposure to violence affects orientation to threat. There is some suggestion that earlier and more chronic exposure might blunt response, after an initial effect of heightening the startle response. For example, Klorman, Cicchetti, Thatcher, and Ison (2003) found such a pattern when they examined electromyograms in response to acoustic startle in maltreated children 3 to 11 years old. The Limbic System The limbic system is a subcortical network regulating emotion, stimulation, and memory. Along with the amygdala, this system processes, interprets, and integrates emotional experience and information (Clugnet & LeDoux, 1990). The system interprets positive versus negative emotional experience, initial organization of an experience affectively and cognitively, and response tendency for socially relevant stimuli. As the seat of emotional memory, it also is implicated in activation of motor behavior and neuroendocrine (stress) response. Within this

system, the hippocampus is potentially affected by violence exposure because it is involved in storage of sensory information and in stress activation (Phillips & LeDoux, 1992). Threat may affect the nature of emotional storage as well as the efficiency. Studies of abused children suggest no difference or only marginal effect in hippocampal volume among maltreated children with PTSD (De Bellis, 2005). Hippocampal volume has not been shown to relate to violence (De Bellis, 2005). However, as hippocampal atrophy relates to impairment in memory functions and is implicated in HPA axis dysregulation, some suggest it may be an important effect through which violence exposure contributes to psychopathology (Mead et al., 2010). This view emphasizes perceptual and response bias over impulsive responding in understanding how community violence might affect psychopathology and behavior. Cortical and Subcortical Areas In concert with the limbic system, cortical areas of the brain serve to modulate, organize, and moderate emotional and behavioral impulses for adaptation and goal-directed activity (Perry, 2008). Since the cortex takes the longest to develop and is the site of complex decision making, it is thought that violence impacts cortical physiology in part through harm to lower areas that affect the “input.” Also, severe challenges, such as environmentally harsh or under- or overstimulating experiences, might impair development of cortical areas that are the neural substrates of important executive functions (Cicchetti & Toth, 2005; Perry, 2008). Perry (2008) reported that rats raised in an environmentally enriched setting have 30% higher synaptic density in the cortex than rats that were raised in an environmentally deprived setting and suggested that this effect process is analogous to the effects of exposure to violence and neglect. Magnetic resonance imaging and other brain scans in children exposed to abuse and violence have shown differences in white and gray matter volumes and gray matter thickness in cortical areas (e.g., Choi, Bumseok, Rohan, Polcari, & Teicher, 2010; De Bellis, 2005). Mead et al. (2010) provided an extensive review of diverse studies that might together provide a basis for formulating the neurodevelopmental impact of violence exposure, particularly family violence victimization, on cortical development. Although the majority of studies focus on abused and/or neglected children, some focus on media violence exposure. This latter focus may be informative as it is a form of violence exposure that has similarities to community violence. Media violence is not direct victimization but instead is tantamount to being in a violent and violence-accepting environment or as a witness. Also, the violence exposure is not due to perpetration by a caregiver. That the findings seem consistent across media- and abusebased exposure lends some credibility to looking for similar effects due to community violence exposure.

Neurotransmitter Systems Studies of violence in relation to psychopathology have focused mostly on three neurotransmitter systems: dopamine (DA), norepinephrine (INN), and serotonin (5-HT). The DA system has been implicated through its linkage to differences in novelty seeking, reward responding, approach motivation, and emotion; and through the cortical effects on cognition, working memory, self-regulation, and planning. Thus, much of the work on understanding the

role of the DA system in psychopathology has been related to externalizing problems, such as attention-deficit/hyperactivity and conduct disorder. The theoretical linkage from violence exposure to these empirical findings is that substantial and sustained exposure could induce long-term downregulation of central DA activity (Mead et al., 2010). Behaviorally, this would translate to more impulsive and aggressive behavior with related changes in motivation, cognitive processes of aggression, and self-control. Elevated INN has been related to trauma symptoms and violence exposure in numerous studies (De Bellis, 2005). It has been shown to differentiate children with such symptoms and history from those without such history. It also has been shown to relate to duration or trauma and predicted symptom levels posttrauma (De Bellis, 2005; Delahanty, Nugent, Christopher, & Walsh, 2005). INN is one of the few neurotransmitters studied in relation to exposure to community violence (Wilson et al., 2002). In that study, victimization that was related to community violence predicted INN level for both males and females. There was a significant relation between witnessing or hearing about violence and INN level across the day for males but not females. Although the study used a small sample (n = 56) of a single ethnic group (African American), it is one of the few studies to document neurotransmitter effects related to violence exposure. Since selection screening was for healthy youth, members of this sample probably had little or no abuse histories, so these results may represent a linkage of findings about community violence empirically to those observed in studies of maltreated children. Across studies, including Wilson et al. (2002), the primary hypothesis has been that INN functioning is changed by stress exposure and that the effect is to lower the behavioral inhibition system, which may not lead to more impulsive and aggressive behavior. Additionally, it is thought that this exposure will reduce aversion to violence through blunting the stress response. There has been substantial documentation of the relation of 5-HT depletion and impulsive and aggressive behavior (Duke, Bègue, Bell, & Eisenlohr-Moul, 2013; Pine et al., 1997). Depletion of 5-HT has been linked to lessening of harm avoidance and to greater likelihood of anxiety and depression. Animal studies indicate that regulation of the 5-HT system is affected by the animal's early rearing environment, particularly abusive rearing environments. These patterns are taken to imply that 5-HT functioning can be affected by violence exposure and may be one way of affecting brain development and functioning. As with INN and DA systems, there is no compelling evidence of what effects, if any, community violence has on the 5-HT system. However, there is substantial indication of the relation of functioning of these systems and violence exposure or psychopathology. Thus, there is good reason to investigate how these might be affected by community violence.

Cortisol and Cortisol Regulation The area of brain functioning that has received the most attention in relation to violence exposure, particularly related to maltreatment, has been cortisol regulation and the HPA axis (Cicchetti & Rogosch, 2001; Gunnar & Vazquez, 2006). The primary hypothesis is that stress, particularly traumatic or overwhelming stress, leads to increased cortisol levels and/or impaired regulation of cortisol levels such that reactivity patterns (more reactive, longer to

return to baseline) and diurnal patterns are affected. During acute stress, the HPA axis is activated and cortisol serves to downregulate hormonal secretions and to stop many metabolic and neuronal defensive and immune reactions in order to mobilize coping with stress. Thus, cortisol as a catabolic hormone activates energy mobilization, focused attention, vigilance, and memory formation. The effect is to distort emotions related to threat and cortical processes essential in processing, evaluating, and managing threat (Gunnar & Vazquez, 2006). Ongoing and particularly shocking stress (such as violence exposure) is seen as potentially permanently affecting this system, perverting brain functioning, and perhaps affecting physiology. The behavioral effect is threat-focused anxiety, depression, and/or aggression (Twardosz & Lutzker, 2010). Although a large number of studies have marked a relation among cortisol level, reactivity, and/or diurnal patterns and violence exposure, the results have varied across studies. Sometimes variations show complex patterns that vary by type of maltreatment or violence exposure, gender, and which cortisol regulation characteristic was implicated (Cicchetti & Rogosch, 2001). Also, in many studies, it is difficult to disentangle what might be psychopathology basis for cortisol regulation differences and what is attributable to violence exposure that might also explain psychopathology occurrence. As with other areas of neurodevelopmental violence effects, the vast majority of studies focus on maltreatment as the source of violence exposure, often sampling children with severe and chronic abuse histories. The next largest set of studies in this area did not differentiate source of violence exposure. These latter studies often summed up types of violence experiences to relate overall exposure. One exception was a study that showed negative relations for both peer victimization and witnessing violence and basal cortisol levels among children living within urban neighborhoods with high rates of violence (Kliewer, 2006). These effects were found after controlling for stressful life events, gender, and age. Recently Aiyer, Heinze, Miller, Stoddard, and Zimmerman (2014) examined the longitudinal relation of cumulative violence exposure of an African American sample residing in a low-income, high-crime community. A cumulative exposure to violence effect on cortisol was found for males, with those exposed having a more attenuated response pattern. Notably, the effect of cumulative exposure to violence on cortisol was moderated by the presence of fathers' support during adolescence. This study did not limit focus to community violence, but it was part of the cumulative violence exposure index. The relation of cortisol patterns and psychopathology among violence-exposed children has been stronger for depression or internalizing disorders, with the relation primarily to higher levels of cortisol (Cichetti & Rogosch, 2001; Mead et al., 2010). Lower levels have shown relation to aggressive tendencies among maltreated and nonmaltreated children, albeit inconsistently and with gender variations in any effect and type of effect (Cicchetti & Rogosch, 2001; Gunnar & Vazquez, 2006). Also, there is evidence that cortisol regulation effects related to abuse may be influenced by entry into nonabusive care and may be responsive to targeted intervention (Cicchetti, Rogosch, Toth, & Sturge-Apple, 2011; Fisher, Gunnar, Chamberlain, & Reid, 2000). Other studies have shown long-term effects of violence exposure on cortisol regulation (e.g., Peckins, Dockray, Eckenrode, Heaton, & Susman, 2012). However, the promising findings in some studies show a pattern of inconsistencies across studies that have

raised questions about how much cortisol patterns may reflect the current situation more than developmental effects of prior experience (see Gunnar & Vazquez, 2006, for an excellent review of the issues). Moreover, how community violence exposure may compromise cortisol regulation in a manner that could reasonably be interpreted as contributing to psychopathology remains relatively unexplored. In a study not of violence exposure but of “neighborhood disorder related to residence in impoverished urban communities,” Dulin-Keita, Casazza, Fernandez, Goran, and Gower (2010) reported that neighborhood disorder was predictive of lower serum cortisol levels. The number of years spent living in poverty related positively to overnight cortisol level and extent of dysregulation of cardiovascular response (i.e., muted reactivity). A similar finding was reported by Evans and Kim (2007), which tracked poverty status, cumulative life stress, and health and reactivity to an acute stressor. The cumulative stress measure appears quite similar to the neighborhood disorder indicators used by Dulin-Keita et al. (2010). Evans and Kim reported that the more years one lived in poverty, the more elevated overnight cortisol was and the more poorly regulated was the cardiovascular response (i.e., muted reactivity). These studies suggest that studying the effects of community violence on cortisol regulation would be very useful.

Arousal/Resting Heart Rate As violence and other stressful experiences typically increases arousal, one presumed effect of exposure to community violence in neurophysiology is the effect on resting heart rate (Perry, 2008). Indeed, studies have documented both increased arousal and hypoarousal of youth exposed to violence, although most of these studies have been of specific clinical populations and with a focus on trauma (Margolin & Gordis, 2000). A few studies have examined the link between exposure to community violence and arousal, and conclusions have been inconsistent. For example, high exposure to community violence has been associated with a lower resting heart rate in a sample of urban school-age children (Krenichyn, Saegert, & Evans, 2001). Similarly, Cooley-Quille, Boyd, Frantz, and Walsh (2001) reported adolescents with higher levels of community violence exposure had lower baseline heart rates than less exposed subjects in response to media violence. In contrast, Wilson et al. (2002) reported higher heart rates in exposed youth. Cardiovascular recovery was not affected by duration of poverty exposure. Notably current poverty status did not relate to these outcomes. These studies suggest the value of consideration of arousal level muting as one effect of community violence exposure. However, these studies also raise the question of confounding of effects with residence in impoverished communities, particularly those marked by more “disorder” or more limited structural resources for social organization (M. Wilson, 1988).

Implications for Understanding and Affecting Community Violence Exposure This broad set of neurodevelopmentally focused studies has identified many promising areas of research on how community violence exposure affects youth, including some indicators of what

processes might be implicated in psychopathology. The applicability of each and every area of study remains highly uncertain at this point simply due to the lack of more than a few studies of community violence effects as a specific risk factor. The complexities of differentiating such exposure, isolating effects of community violence exposure from frequently co-occurring developmental risk factors, and the size and scope of studies needed to explore neurodevelopmental effects of such exposure competently are all impediments to such efforts. However, this important prior work has suggested many promising topics for research and avenues for likely effects. Among these is how genetic variation in susceptibility to violence exposure and disordered and neglectful/non-nurturant developmental environments affect brain physiology, neurotransmitters, and important neurophysiological systems. Others are to track how variation in family management practices and/or neighborhood social processes affects violence exposure and the extent of the effects on psychopathology of those exposed. Similarly, following patterns of stress reactivity and impulsivity subsequent to incidents of violence exposure or over time due to chronic community violence exposure would help illuminate how this form of violence exposure compares to other forms.

Interventions for Community Violence Exposure As with understanding other aspects of the relation of community violence exposure, there are few direct tests of interventions meant either to lessen community violence exposure or to mitigate impact. Most efforts have focused on reducing violence exhibition by youth or by those directly involved in their supervision (e.g., reduce use of violence by parents). Although victimization was also considered in some cases, most efforts to affect community violence levels only measure impact on perpetration of crimes. Moreover, most reports are descriptive only or use empirical comparisons of limited methodological quality, which limits confidence in the validity and/or robustness of the findings. Thus, efforts to reduce youth crime and neighborhood-level violence can be considered potentially valuable in reducing community violence exposure. Similarly, few studies have focused on mitigating exposure to community violence specifically. Often violence exposure measurement does not differentiate source of violence exposure (e.g., community versus familial) or type of exposure (e.g., direct victimization from witnessing). A set of studies has approached mitigation of violence effects, including community violence as a traumatic experience (Lieberman & DeMartino, 2006). These studies treat the exposed youth as traumatized and typically apply individual or family interventions to help reduce the trauma symptoms and the resulting behavioral effects. Also, some studies focus on aggression or conduct problems as the targeted effect of community violence exposure, although in most cases it is not clear how exposure and involvement in violence as a perpetrator comingle in the effects (Tolan, 2001b).

Efforts to Reduce Community Violence Community Organization for Violence Deterrence

One strategy for reducing community violence that has gained much attention is helping communities to organize, be vigilant, and act to make safe areas for children. The intent is to engage residents to help monitor their own communities, notifying police if violence seems imminent and thereby decreasing opportunities for violence. Often these efforts enlist joint efforts by police, child welfare, churches, other civic organizations, and youth services. One variant on this approach is to organize for communication and intercession to block youth violence, particularly gang-related violence that may be brewing over time (Braga & Weisburd, 2012). Often in this variant, an agency has staff that engages with gang-related youth who are informed about when disputes might be leading to violence and can stave off conflict escalation. In addition, there is a concentration of legal prosecution efforts on violence, particularly shooting, such that resorting to such violence is more likely to lead to prosecution and incarceration. A third element is to positively affect police—neighborhood relations, educating youth and minimizing any broad harassment or intimidation efforts (Graczyk & Tolan, 2005). Not infrequently, these efforts incorporate block meetings by police with residents and other aspects of community policing that can affect crime rates (Braga & Weisburd, 2012). Although there are no reports of effects on reducing exposure to community violence, there is evidence of effects on community violence overall from a meta-analysis (Braga & Weisburd, 2012) and specific tests of programs within this general approach (Braga, Kennedy, Waring, & Piehl, 2001). These effects are small. Within multisite studies, it is common to find effects in some but not all of the included sites. Overall, it appears that neighborhood organization could be useful in reducing youth exposure to community violence. Neighborhood Watch Programs Neighborhood watch programs fall within the category of neighborhood organization, but they limit citizen organization to monitoring neighborhood settings. Public notification of such monitoring in an effort to discourage violence and crime has been tested most frequently. In addition, these programs often emphasize norms of pride about the community upkeep and safety and are thought to improve neighborhood engagement of citizens. Often these efforts include “target hardening,” meaning making citizens aware of how to be less easily victimized. Bennett, Holloway, and Farrington (2008) conducted a meta-analysis of 15 samples across 10 studies of neighborhood watch programs, reporting an average effect of a 16% drop in crime rates. However, the primary contributor to crime rates studied was burglary, not violence. This review noted an average decrease in victimization of 9% from such programs but noted that this was not statistically significant and that only three such tests with adequate evaluation design could be found. Of those, two out of three favored the control condition. These findings suggest that neighborhood watch may deter certain types of crime but may have little impact on exposure to violence. However, any conclusion is quite preliminary without an adequately conducted test of impact on community violence exposure. Community Policing A related effort to the previous two community organization approaches is to alter policing practices from specialists investigating and solving specific crimes to a community presence in

order to discourage crime and identify methods for prevention. Police are assigned to the overall responsibility for a neighborhood's crime and the residents' safety and security. This method has been characterized as a shift of policing from crime solving to crime prevention. The best efforts focus on working with community members to learn about the community, create communication channels with the police that build trust and increase security, and work to promote safety and crime-reducing norms and activities at the neighborhood level. Often a local office is set up for police contact. It is thought that a greater connection between community members and police increases understanding and communication, which can lead to diffusing emerging violence and should lead to more effective police strategies for addressing crime causes and effects, which in turn should prevent crimes, including violence (Graczyk & Tolan, 2005). Careful studies of community policing have shown effects on perceived order and safety in the community, and one has documented reduced victimization (Wycoff & Skogan, 1993). However, a recent meta-analysis of 65 programs across multiple countries reported a positive but not statistically significant effect on crime (5%–10%) (Gill, Weisburd, Telep, Vitter, & Bennet, 2014). Similarly, there was a small but not statistically significant effect for felt safety. It should be noted that there was considerable variation in effect sizes across the included studies and there was considerable variation in what comprised the different programs. The implication may be that the way community policing is done and, in particular, its level of implementation can relate to effects. When the primary principles are followed and multiple components are included, effects can be found for violence levels in communities and probably also for victimization. However, the way in which community policing affects children's exposure to community violence is still unsubstantiated. Gun Control and Management Efforts There has been considerable interest in the role of access and use of firearms as a contributor to youth exposure to community violence. In part this is because of the prominence of firearms, often illegally obtained, in community violence among youth (National Research Council of the National Academies, 2004, 2013). It is also due to the corresponding high rates of youth violence, particularly gang-related violence, in these communities. Although the value of gun control remains one of the most contentious issues in our society and within academic communities, research has suggested some findings relevant to reducing community violence that may be important in efforts to reduce exposure to such violence. These efforts can be categorized into four groups. The first is community-focused efforts to educate about the relation between gun access and lethality of violence, safe storage and proper use of weapons, and efforts to lessen youth access to guns legally or illegally. These efforts also can include community organization to affect norms about gun use, gun violence, and community responsibility for monitoring crime in the community à la community policing strategies. A second community approach is to try to obtain guns owned by youth and others through amnesty for turning in illegal firearms or buy-back programs. A third strategy is for legislation to increase penalties for use of guns in crimes, for illegal sale or buying of guns, to limit the number and type of guns that can be purchased during a given time period, and to require background checks. Fourth, law enforcement campaigns can aim to reduce gun violence. These campaigns can include concentrating on the policing of hot spots (areas of particular interest to

law enforcement because of a concentration of criminal activity) to reduce opportunity for gun violence to occur, community-police engagement, gun courts that speed up the court processing and trial of crimes committed with firearms, stricter supervision of probation and paroled offenders who utilize guns in committing a crime, and required educational programs for first offenders. A recent meta-analysis was published by Makarios and Pratt (2012) of 29 studies categorized into these four types of interventions to determine the effects of policy and community interventions to reduce gun violence. Overall, they found a small but statistically significant decrease in violent gun crime (r = –0.14). When examining more specific interventions, small reductions on average were reported for gun laws (r = –0.09) and a significant moderate effect for law enforcement campaigns (r = –0.23). More specific analysis of gun laws found that weapons bans had significant, moderate effects on reducing gun crime (r = –0.19). Policing strategies in general had an effect size equivalent to law enforcement campaigns (r = –0.23). A similar effect was found for community education and engagement programs (r = –0.27). The latter finding may be particularly relevant in reducing violence exposure of youth. This finding suggests that consideration of efforts to reduce gun violence should include such community engagement strategies and that they are most useful also in affecting exposure because they include elements with some evidence of reducing exposure.

Efforts to Reduce Effects of Exposure The scholarly literature contains many descriptions of efforts designed to reduce exposure or mitigate its impact on youth but few reports of outcomes (e.g., Bowen, Gwiasda, & Brown, 2004; Horowitz et al., 2005). Recently, two large initiatives were launched as federally funded and coordinated demonstration efforts, intended to permit testing of effects. The Office of Juvenile Justice and Delinquency Prevention launched Safe Start Promising Approaches, which funded efforts in 15 sites to develop and/or implement programs meant to mitigate effects on trauma and other problems, although decreasing exposure was also part of the charge (Jaycox et al., 2011). The consistent goal was to ameliorate the negative impacts of violence on children ages zero to 12. Programs provided differed by sites. Also, some community efforts focused on the development of specific services while others attempted to organize or support existing services. Some provided services to children only while others served both children and their families or primary caregivers. The programs also varied within site and in how recipients of services were identified. Most targeted various exposure levels of domestic violence, with some targeting families with child abuse, but others included exposure to community violence. Launched in 1999, an evaluation report was recently released (Jaycox et al., 2011). Most notably, the report noted that outcome effects were often difficult to ascertain with any certainty, which precluded conclusions of effects across the initiative. Nor were specific effects for any given site presented. Underenrollment and attrition issues, evaluation design limitations, inadequate evaluation resources, and a premature end of the evaluation marked the projects. The retained sample in programs at most sites was too small to permit adequate power to detect expectable effects. For example, some sites had enrollment rates as low as 2% of targeted participants. Similarly, the Attorney General's National Task Force on Children Exposed to Violence report

led to a national initiative called Defending Childhood. In this case, eight cities were provided funding and technical support for community organizing to provide services to reduce children's exposure to violence (not just community violence) and to make services available that were thought to lessen the impact of such exposure. Most of the funded cities and Native American groups (included in the funding) focused on getting family-focused services to children exposed to domestic violence or abuse. However, the intent was to also address those exposed to community violence. This initiative, like Safe Start, included a variety of efforts. Therefore, surmising what might be effective was difficult since each individual effort had substantial methodological limitations (Swaner & Kohn, 2011). The Defending Childhood initiative is still in progress. However, interim reports suggest major impediments to implementation of programs as intended and engagement of families that are consistent with those in the evaluation report about Safe Start Promising Approaches (Jaycox et al., 2011). Thus, this experience in implementing needed programs provides an unfortunate but instructive example of how difficult it can be to carry out community-based violence-mitigating efforts. Conceptual organization (what is thought to be needed), implementation challenges (getting needed efforts in place and utilized), and adequate methodological quality to permit inference are all necessary to complete adequate evaluation to determine what efforts might mitigate exposure or its impact. Each of these factors can be a major challenge. Together, they can overwhelm good intentions. It should be noted that these elements often involve large-scale community engagement efforts, shifting of some long-held approaches to services violence prevention. At the same time, identifying viable comparison conditions or communities can be difficult; and implementation in a sufficient number of neighborhoods for valid statistical inference can overwhelm budgets and the operational capabilities of evaluators. The lesson may be to consider some smaller-scale, more controlled demonstrations to establish efficacy. Furthermore, basing work on collaboration models that have shown benefits for other youth problems could then be implemented, marrying promising programming with promising organization (Hawkins, Oesterle, Brown, Abbott, & Catalano, 2014). Also, it is important to focus on community violence exposure apart from family violence as integral to the evaluation. This could be via a focus on community violence exposure specifically or as an attempt to parse the relative level of each among participants and how those levels relate to effects. Police and Social Worker Collaboration Marans and Berkman (2006) developed a variant on the community policing approach as a way to get services to youth exposed to violence that differs from community organization. They formed a collaboration between community police and mental health professionals to address effects of witnessing violence on children. The approach combined extensive training of the police in child development theory and violence effects on children and direct intervention by mental health caseworkers as soon as children had witnessed a violent crime. Marans and Berkman noted several promising effects on police approach and on the handling of domestic violence, but empirical evidence of specific impact on youth exposed to community violence has not been reported to date. As with most other approaches discussed here, there is some suggestion of utility in regard to violence exposure impact. However, that

utility awaits empirical testing. Trauma-Focused Efforts to Mitigate Violence Exposure Effects As with several other approaches to community violence exposure, the literature describes multiple efforts but few target youth exposed to community violence. Or the efforts do not provide results of evaluation to judge effects. There have been multiple trials of psychological interventions for children thought to be traumatized due to maltreatment (Leenarts, Diehle, Doreleijers, Jansma, & Lindauer, 2013). This systematic review and other conceptual reviews suggest that cognitive-behavioral approaches that systematically include a mix of emotional support, opportunity to talk about the trauma in a safe place, anxiety-reducing and thoughtredirecting practices can reduce PTSD symptomology. One example, which included violence exposure not specific to maltreatment, was reported by Stein et al. (2003). They reported on a program entitled Cognitive Behavioral Intervention for Trauma in Schools. The program was designed to reduce PTSD, depression, and anxiety among children with symptoms of PTSD and who were exposed to violence. The 10-session school-based intervention taught cognitivebehavioral skills in a group format, led by mental health professionals, with six to eight students per group, using a mixture of didactic presentation, examples, and games to solidify concepts. Techniques used include relaxation training, combating negative thoughts, reducing avoidance, developing a trauma narrative, and building social problem-solving skills. In addition to the group program, students were provided one to three individual sessions and optional sessions that included parents. Teachers in the school were provided an in-service on trauma, violence, and the intervention. Youth in the program showed significant decreases in trauma symptomology compared to randomized waitlist controls. No effects on problem behavior were found, however. This trial produced results consistent with other tests of this intervention as a clinical treatment for maltreated youth and other similar approaches (Leenart et al., 2013). This single example is a weak basis to infer how community violence exposure might be mitigated through cognitive-behavioral intervention. However, this is among the few efforts meant to mitigate violence exposure in general, not just community violence exposure specifically. This type of approach deserves priority for testing.

Lessons from School-Based Programs School-based prevention has been a popular approach to reduce aggression perpetration. Further, because these efforts are focused on youth violence in a setting where youth spend a large portion of their time, these findings can have two types of implications to affect exposure to community violence. First, these settings provide some attention to the ecology and social basis of violence exposure, with a few considering victimization and other exposure to violence in addition to the main focus on perpetration. Second, while not out in the community, school-based violence is sometimes considered a part of community violence exposure for youth. Therefore, these interventions may provide direction for direct impact on community violence exposure that occurs at school. There is some evidence that some school-based violence prevention programs reduce victimization (Hahn et al., 2007). Similarly, there is some evidence that bullying prevention

efforts also can reduce victimization. In both cases the effects can be from reducing perpetration so fewer students are victims or by making aggression/bullying less acceptable and providing potential victims coping and assertion skills, The best evidence for bullying programs is those that include family-focused components are helpful (Ttofti & Farrington, 2010). For example, Ttofti and Farrington (2010) examined 89 reports and noted an average reduction of victimization of 17% to 20%. The effect sizes varied substantially (from about 1.00 standard deviation to negative effects), with some characteristics relating to effect sizes. They noted that programs that were delivered by or focused on peer groups seem to have lower and sometimes negative effects. Among the effective programs were elements such as clear and explicit display and enforcement of norms in school against violence, engagement of families as part of this norming effort, problem-solving opportunities when conflicts arose, and opportunities for students and staff to increase social and emotional skills. Also, some programs utilized engaging games to help students learn methods to avoid bullying or to help stop victimization. Thus, a variety of approaches had some effects on victimization, at least as measured as bullying, within schools. Notably, the effect sizes tended to be larger for demonstrations in countries other than the United States. Thus, it is not clear how these efforts might translate to reducing community violence or affecting exposure impact and how these might be useful in the United States. Two recent large-scale tests of school-based violence prevention suggested that what might affect victimization differs from what might affect perpetration. The Multisite Violence Prevention Project (2009) tested two programs for middle schools meant to affect violence rates (perpetration and victimization). One program utilized cognitive skills training and norms promotion, similar to what characterizes many programs designed to reduce violence use and acceptance and was applied as a universal classroom-delivered program. The other focused on youth whom teachers identified as among the more aggressive. It utilized small family groups to help with parenting skills, parent-child communication, and group support for nonaggressive problem solving. Testing of effects showed no overall differences for either intervention or a combination of the two for victimization, although some effects of the selective family intervention were found (Multisite Violence Prevention Project, 2009, 2014). There was some suggestion of lower victimization of the combined program for those students with very elevated aggression. Similarly, Espelage, Low, Polanin, and Brown (2013) reported that test of a curriculum, Second Step, that focused on many of the same cognitive features as the Multisite Violence Prevention Project (and many other such programs) showed no effects on victimization even though there were positive effects for perpetration of bullying. As with other tests of in-school programs, both of these efforts focused on withinschool victimization, so results may not transfer to tests that measure effects of community violence.

Advancing Knowledge, Practice, and Policies Related to Community Violence Exposure

Violence exposure has harmful effects on development. Numerous studies have implicated exposure to violence as contributing to risk for psychopathology symptoms and disorders and in exacerbating psychopathology. Community violence, one form of violence exposure, has been studied less than other forms, but research suggests its impact is similarly deleterious. In addition, it probably affects psychopathology through comparable psychological and biological processes. Community violence differs from other forms of violence in that it occurs outside the home and is not due to violence between family and friends. It also differs from these latter forms of violence in that it may affect youth by virtue of residing within a geographic region. Thus, its forms can vary from isolated horrible events to an atmosphere where violence is endemic and frequent. Most specifically, it can range from direct victimization to witnessing victimization of others to being aware of violence nearby or occurring to neighbors. Also, unlike family violence (including partner violence, child abuse, and sibling violence), community violence tends to be concentrated in particular neighborhoods, particularly urban impoverished communities. Thus, it is substantially related to economic class and location, which is not the case for other forms of violence. This makes community violence a political and social policy issue as much as psychological and psychopathological issues. Over the past 20 years, considerable theory, policy analyses and suggestions, and empirical study have been applied to community violence. In addition, considerable effort and advances have occurred in understanding the importance of the definition of violence applied, how distinct measurement of exposure types and sources are, and how these are interwoven into developmental tracking and intervention studies. This collective work has revealed important understanding of the issues, including alarming prevalence among some children in particular communities. The starkness of these findings and the concentration of violence within poorer communities in larger urban centers suggest the importance of contextualized understanding of impact. This can include better understanding of what community-level actions can reduce violence and exposure but also what might be needed to mitigate harm from exposure in communities marked by limited social and economic resources and limited political importance. There is sufficient evidence that community organization, community policing, and community awareness can reduce crime rates and probably reduce community violence rates. Therefore, such actions deserve to be central in efforts to reduce community violence exposure and to the study of what makes a difference. It has also shown that while effects of violence exposure can be traumatic to some individuals, this is not universally so. In fact, there is mixed evidence for any form of violence exposure affecting symptoms of PTSD, with thin evidence of effects extending to disorder incidence. Evidence is similarly strong for violence exacerbating rather than affecting conduct disorder symptoms and other externalizing disorders and depression. Evidence is less robust for violence affecting anxiety symptoms apart from PTSD, but is the available evidence shows effects consistent with those shown for other symptoms and other disorders, suggesting potential effects for some exposed. As expected, the few studies that have examined the relation between psychopathology and impact of subsequent exposure to violence have demonstrated the transactional quality of the interdependence. Those with mental health

problems seem most susceptible to exposure harm. In some cases, particularly for those with aggressive tendencies, the exposure may be due to those tendencies with additional increased impact in the presence of such psychopathology. Similarly, when related to different forms of violence, those with prior abuse or exposure to domestic violence are most affected by community violence. Impact is also dependent on multiple personal and familial characteristics, suggesting that those growing up with fewer resources or poorer functioning are also most susceptible to violence risk in comparison to others residing in similar communities. The accumulated research on community violence effects on psychopathology underscores the transactional nature of its development with attention to group level as well as individual differences in exposure risk and impact processes. Although the components of models that can explain and provide guidance for additional studies can be found significant gaps still exits in the empirical basis for the link between community violence and psychopathology occurrence, course, and mitigation. When theories of processes are applied, whether framed within a neurodevelopmental approach or as psychological processes of emotion and cognition, a common element is strong dependence on theories and studies from maltreatment exposure and explanations of violence or aggression perpetration. This is due to limited research on community violence, limited differentiation of community violence from other forms of violence when it is studied, the larger and more extensive set of studies on maltreatment and violent behavior, and the likelihood that there are similar if not shared processes of impact. A more carefully drawn definition of violence is critical to permit confident understanding of the meaning of various studies. But violence also must be differentiated into levels of seriousness, sources, and forms. Doing this will allow us to determine whether effects and processes are similar across community and family violence and are related to differences in level of exposure to and directness of involvement in the violence. In addition, more studies are needed that test for community violence exposure in particular as a source of stress and as impeding self-control. Also, it is necessary to trace neurobiological processes, test for differences in physiology, and document the stability/transience of effects on those processes thought to underlie/characterize psychopathology. In reviewing areas related to community violence, much of what we know is based on inference from studies of exposure to violence in general or due to other forms. There are still basic questions of how community violence exposure specifically relates to likely symptoms and behavior change. While there is great value in testing the application of lessons from these findings about violence in general to exposure to community violence in particular, this application needs to take into account differences in how this type of exposure occurs and in likely co-occurring risk factors. Interventions for community violence exposure that are designed to test theories of psychopathology as well as those intended to demonstrate the potential for public health benefits if taken to scale are both needed, and should be priorities for future research. There have been a notable number of failures to implement adequately or to complete study of community violence interventions. Therefore, there is very limited scientific information about impact of any approach and even scarcer basis for judging one approach as advantageous over others. In part, this state of knowledge can be attributed to the scope and complexity of

implementing interventions that attempt to address structural issues such as gang and criminal activity in a community, or operate with limited police availability and engagement with the community, and/or attempt to address only violence when it is part of a collective concentration of social problems (Tolan & Guerra, 1998; M. Wilson, 1988). It may be important to focus on how communities can become more engaged to form and express collective efficacy (Sampson et al., 1997). Similarly, it may be advantageous to explore how families and youth can be reached through greater school-community engagement or encouraging neighborhood supports (Tolan et al., 2003). If affecting the social forces that seem to undergird rates of violence and therefore likely violence exposure in communities is beyond the reach of most policy or service efforts, perhaps some of the promising family support, safe space, youth engagement, and youth monitoring efforts that show promise from prior trials could be utilized to explicitly limit exposure and mitigate harm from exposure.

Conclusion Community violence exposure remains a compelling social and psychological problem that deserves more of the careful and informative scientific work extending what has been done in the past 25 years. In particular, what is needed are: (a) more carefully defined samples; (b) measures of community violence exposure that delve into extent of exposure (e.g., direct victimization versus witness versus being aware of) and of exposure with considerations of other bases for violence (exposure, chronicity of exposure); and (c) specific theoretically driven formulations of the relation of community violence exposure to context, family, individual behavioral, affective, and cognitive features, and neurodevelopmental underpinnings; and (d) analytic approaches that can trace causal influence of community violence exposure to changes in psychopathology symptoms and conditions. With such design features, it is plausible that knowledge could be accumulated to move us from the basic understanding available at this point toward delineation of the basis, effects, processes of impact, and needed interventions and policies for community violence exposure. The field seems poised to build on prior work to incorporate more careful definitions of key constructs and to apply methods and findings from adjacent studies of violence, aggression, crime prevention, and neighborhood organization, and on processes of developmental psychopathology. With such careful and substantial attention, understanding of community violence could be greatly improved in a relatively brief time and could be used to help lessen this harmful part of the lives of too many children.

References Aiyer, S. M., Heinze, J. E., Miller, A. L., Stoddard, S. A., & Zimmerman, M. A. (2014). Exposure to violence predicting cortisol response during adolescence and early adulthood: Understanding moderating factors. Journal of Youth and Adolescence, 43, 1066–1079. doi: 10.1007/s10964–014–0097–8. American Psychological Association Presidential Task Force on Violence and the Family.

(1996). Violence and the family: Report of the American Psychological Association Presidential Task Force on violence and the family. Washington, DC: American Psychological Association. Attar, B. K., Guerra, N. G., & Tolan, P. H. (1994). Neighborhood disadvantage, stressful life events, and adjustment in urban elementary-school children. Journal of Clinical Child Psychology, 23(4), 391–400. doi: 10.1207/s15374424jccp2304_5 Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135(6), 885–908.doi: 10.1037/a0017376 Bennett, T., Holloway, K., & Farrington, D. (2008). The effectiveness of Neighborhood Watch. Campbell Systematic Reviews, 18, 1–51. Bentley, K. L., Thomas, D. E., & Stevenson, H. C. (2013). Raising consciousness: Promoting healthy coping among African American boys at school. In C. Clauss-Ehlers, Z. Serpell, & M. Weist (Eds.), Handbook of culturally responsive school mental health: Advancing research, training, practice, and policy (pp. 121–133). New York, NY: Springer. Bowen, L. K., Gwiasda, V., & Brown, M. M. (2004). Engaging community residents to prevent violence. Journal of Interpersonal Violence, 19(3), 356–367. Bowen, N. K., & Bowen, G. L. (1999). Effects of crime and violence in neighborhoods and schools on the school behavior and performance of adolescents. Journal of Adolescent Research, 14(3), 319–342. doi: 10.1177/0743558499143003 Boyd, R., Cooley, M., Lambert, S., & Ialongo, N. (2003). First-grade risk behaviors for community violence exposure in middle school. Journal of Community Psychology, 31, 297– 314. doi: 10.1002/jcop.10047 Bradshaw, C. P., Rodgers, C., Ghandour, L., & Garbarino, J. (2009). Social-cognitive mediators of the association between community violence exposure and aggressive behavior. School Psychology Quarterly, 24(3), 199–210. doi: 10.1037/a0017362 Braga, A. A., Kennedy, D. M., Waring, E. J., & Piehl, A. M. (2001). Problem-oriented policing, deterrence, and youth violence: An evaluation of Boston's Operation Ceasefire. Journal of Research in Crime and Delinquency, 38, 195–226. doi: 10.1177/0022427801038003001 Braga, A. A., & Weisburd, D. L. (2012). The effects of “pulling levers” focused deterrence strategies on crime. Campbell Systematic Reviews. doi: 10.4073/csr.2012.6 Browning, C. R., Gardner, M., Maimon, D., Brooks-Gunn, J. (2014). Collective efficacy and the contingent consequences of exposure to life-threatening violence. Developmental Psychology, 50(7), 1878–1890. doi: 10.1037/a0036767 Bursik, R. J. (1988). Social disorganization and theories of crime and delinquency: Problems

and prospects. Criminology, 26(4), 519–551. Byrne, J., & Sampson, R. J. (1986). Key issues in the ecology of crime. In J. Byrne & R. Sampson (Eds.). The social ecology of crime (pp. 1–22). New York, NY: Springer-Verlag. Caspi, A., McClay, J., Moffitt, T. E., Mill, J., Martin, J., Craig, I. W.,… Poulton, R. (2002). Role of genotype in the cycle of violence in maltreated children. Science, 297(5582), 851– 854. doi: 10.1126/science.1072290 Castro-Alamancos, M., & Connors, B. W. (1996). Short-term plasticity of a thalamocortical pathway dynamically modulated by behavioral state. Science, 272(5259), 274–277. doi: 10.1126/science.272.5259.274 Centers for Disease Control and Prevention. (2014). Youth risk behavior surveillance-United States, 2013. Mortality and Morbidity Weekly Report, 63, 1–48. Chalk, R., & King, P. A. (1998). Violence in families: Assessing prevention and treatment programs. Washington, DC: National Academies Press. Chaux, E., Arboleda, J., & Rincon, C. (2012). Community violence and reactive and proactive aggression: The mediating role of cognitive and emotional variables. Revista Colombiana De Psicologia, 21, 233–251. Choi, J., Bumseok, J., Rohan, M. S., Polcari, A. M., & Teicher, M. H. (2010). Preliminary evidence for white matter tract abnormalities in young adults exposed to parental verbal abuse. Biological Psychiatry, 65, 227–234. doi: 10.1016/j.biopsych.2008.06.022 Chung, H. L. & Steinberg (2006). Relations between neighborhood factors, parenting behaviors, peer deviance, and delinquency among serous juvenile offenders. Developmental Psychology, 42, 319–331. doi: 10.1037/0012-1649.42.2.319 Cicchetti, D., & Curtis, W. J. (2006). The developing brain and neural plasticity: Implications for normality, psychopathology, and resilience. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology, Vol. 2: Developmental neuroscience (2nd ed., pp. 1–64). Hoboken, NJ: Wiley. Cicchetti, D., & Rogosch, F. A. (2001). The impact of child maltreatment and psychopathology on neuroendocrine functioning. Development & Psychopathology, 13(4), 783–804. Cicchetti, D., Rogosch, F. A., & Sturge-Apple, M. L. (2007). Interactions of child maltreatment and serotonin transporter and monoamine oxidase A polymorphisms: Depressive symptomatology among adolescents from low socioeconomic status backgrounds. Development & Psychopathology, 19(4), 1161–1180. doi: 10.1017/S0954579407000600 Cicchetti, D., Rogosch, F., Toth, S. L., & Sturge-Apple, M. L. (2011). Normalizing the development of cortisol regulation in maltreated infants through preventive interventions. Development and Psychopathology, 23, 789–800.

Cicchetti, D., & Toth, S. L. (1995). A developmental psychopathology perspective on child abuse and neglect. Journal of American Academy of Child and Adolescent Psychiatry, 34(5), 541–565. doi: 10.1097/00004583–199505000–00008 Cicchetti, D., & Toth, S. L. (2005). Child maltreatment. Annual Review of Clinical Psychology, 1(1), 409–438. doi: 10.1146/annurev.clinpsy.1.102803.144029 Clugnet, M. C., & LeDoux, J. E. (1990). Synaptic plasticity in fear conditioning circuits: Induction of LTP in the lateral nucleus of the amygdala by stimulation of the medial geniculate body. Journal of Neuroscience: The Official Journal of the Society for Neuroscience. Cooley-Quille, M., Boyd, R. C., Frantz, F., & Walsh, J. (2001). Emotional and behavioral impact of exposure to community violence in inner-city adolescents. Journal of Community Psychology, 30, 199–206. doi: 10.1207/S15374424JCCP3002_7 Covey, H. C., Menard S., & Franzese, R. J. (2013). Effects of adolescent physical abuse, exposure to neighborhood violence, and witnessing parental violence on adult socioeconomic status. Child Maltreatment, 18(20), 85–97. Epub 2013 Feb 17. doi: 10.1177/1077559513477914 Crockett, M. J., Clark, L., Hauser, M. D., & Robbins, T. W. (2010). Serotonin selectively influences moral judgment and behavior through effects on harm aversion. Proceedings of the National Academy of Sciences, 107(40), 17433–17438. doi: 10.1073/pnas.1009396107 Currie, J., & Widom, C. S., (2010). Long-term consequences of child abuse and neglect on adult economic well-being. Child Maltreatment, 15(2), 111–120. doi: 10.1177/1077559509355316 De Bellis, M. D. (2005). The psychobiology of neglect. Child Maltreatment, 10(2), 150–172. doi: 10.1177/1077559505275116 De Bellis, M. D., Hooper, S. R., Spratt, E. G., Woolley, D. P. (2009). Neuropsychological findings in childhood neglect and their relationships to pediatric PTSD. Journal of the International Neuropsychological Society, 15(6), 868–878. doi: 10.1017/S1355617709990464 Delahanty, D. L., Nugent, N. R., Christopher, N. C., & Walsh, M. (2005). Initial urinary epinephrine and cortisol levels predict acute PTSD symptoms in child trauma victims. Psychoneuroendocrinology, 30(2), 121–128. doi: 10.1016/j.psyneuen.2004.06.004 Dempsey, M. (2002). Negative coping as mediator in the relation between violence and outcomes: inner-city African American youth. American Journal of Orthopsychiatry, 72(1), 102–109. Dodge, K. A. (1993). Social-cognitive mechanisms in the development of conduct disorder and depression. Annual Review of Psychology, 44, 559–584.

Dodge, K. A. & Somberg, D. R. (1987). Hostile attributional biases among aggressive boys are exacerbated under conditions of threats to the self. Child Development, 58, 213–224 Duke, A., Bègue, L., Bell, R., & Eisenlohr-Moul, T. (2013). Revisiting the serotonin– aggression relation in humans: A meta-analysis. Psychological Bulletin, 139(5), 1148–1172. doi: 10.1037/a0031544 Dulin-Keita, A., Casazza, K., Fernandez, J. R., Goran, M. I., & Gower, B. (2010). Do neighborhoods matter? Neighborhood disorder and long-term trends in serum cortisol levels. Journal of Epidemiology and Community Health, 66, 24–29. doi: 10.1136/jech.2009.092676 Edlynn, E. S., Gaylord-Harden, N. K., Richards, M., & Miller, S. A. (2008). African American inner-city youth exposed to violence: Coping skills as a moderator for anxiety. American Journal of Orthopsychiatry, 78(2), 249–258. doi: 10.1037/a0013948 Eisner, M. (2011). Human evolution, history and violence: An introduction. British Journal of Criminology, 51(3), 473–478. doi: 10.1093/bjc/azr028 Eitle, D., & Turner, R. J. (2002). Exposure to violence and young adult crime: The effects of witnessing violence, traumatic victimization, and stressful life events. Journal of Research in Crime and Delinquency, 39, 214–237. doi: 10.1177/002242780203900204 Elliott, D., & Tolan, P. H. (1999). Youth violence prevention, intervention, and social policy: An overview. In D. Flannery & R. Hoff (Eds.), Youth violence: Prevention, intervention, and social policy (pp. 3–46). Washington, DC: American Psychiatric Association. Espelage, D. L., Low, S., Polanin, J. R., & Brown, E. C. (2013). The impact of a middle school program to reduce aggression, victimization, and sexual violence. Journal of Adolescent Health, 53, 180–186. doi: 10.1016/j.jadohealth.2013.02.021 Evans, G. W., & Kim, P. (2007). Childhood poverty and health cumulative exposure to risk and stress dysregulation. Psychological Science, 18, 953–957. doi: 10.1111/j.1467– 9280.2007.02008.x Evans, S. E., Davies, C., & DiLillo, D., (2008). Exposure to domestic violence: A metaanalysis of child and adolescent outcomes. Aggression and Violent Behavior, 13, 131–140. doi: 10.1016/j.avb.2008.02.005 Fagan, A. A., Wright E. M., & Pinchevsky, G. M. (2014). The protective effects of neighborhood collective efficacy on adolescent substance use and violence following exposure to violence. Journal of Youth and Adolescence, 43, 1498–1512. doi: 10.1007/s10964–013– 0049–8. Fagan, J., & Browne, A. (1994). Violence between spouses and intimates: Physical aggression between women and men in intimate relationships. In A. J. Reiss Jr., & J. A. Roth (Eds.), Understanding and preventing violence, Volume 3: Social Influences (pp. 115–292). Washington, DC: National Academies Press.

Fagan, J., Piper, E., Cheng, Y. T. (1987). Contributions of victimization to delinquency in inner cities. Journal of Criminal Law & Criminology, 78(3), 586–613. Farrell, A., & Bruce, S. (1997). Impact of exposure to community violence on violent behavior and emotional distress among urban adolescents. Journal of Clinical Child Psychology, 26, 2–14. Farrell, A. D., & Sullivan, T. N. (2004). Impact of witnessing violence on growth curves for problem behaviors among early adolescents in urban and rural settings. Journal of Community Psychology, 32, 505–525. Farrington, D., Langan, P., & Tonry, M. (2004). Cross-national studies in crime and punishment. Washington, DC: Bureau of Justice Statistics. Felson, R. B. (1997). Routine activities and involvement in violence as actor, witness, or target. Violence and Victims, 12(3), 209−221. Fincham, D. S., Altes, L. K., Stein, D. J., & Seedat, S. (2009). Posttraumatic stress disorder symptoms in adolescents: risk factors versus resilience moderation. Comprehensive Psychiatry, 50, 193–199. Finkelhor, D., Turner, H. A., Ormrod, R., Hamby, S., & Kracke, K. (2009). Children's exposure to violence: A comprehensive national survey. U.S. Department of Justice, Office of Justice Programs, Office of Juvenile Justice and Delinquency Prevention. Finkelhor, D., Turner, H. A., Shattuck, A., & Hamby, S. L. (2013). Violence, crime, and abuse exposure in a national sample of children and youth: An update. Journal of the American Medical Association Pediatrics, 167(7), 614–621. doi: 10.1001/jamapediatrics.2013.42 Fisher, P. A., Gunnar, M. R., Chamberlain, P., & Reid, J. B. (2000). Preventive intervention for maltreated preschool children: Impact on children's behavior, neuroendocrine activity, and foster parent functioning. Journal of the American Academy of Child & Adolescent Psychiatry, 39(11), 1356–1364. doi: 10.1097/00004583–200011000–00009 Fitzpatrick, K. M., & Boldizar, J. P. (1993). The prevalence and consequences of exposure to violence among African American youth. Journal of the American Academy of Child and Adolescent Psychiatry, 32, 424–430. Foley, D. L., Eaves, L. J., Wormley, B., Silberg, J. L., Maes, H. H., Kuhn, J., & Riley, B. (2004). Childhood adversity, monoamine oxidase a genotype, and risk for conduct disorder. Archives of General Psychiatry, 61(7), 738–744. doi: 10.1001/archpsyc.61.7.738 Fowler, P. J., Tompsett, C. J., Braciszeweski, J. M., Jacques-Tiura, A. J., & Baltes, B. B. (2009). Community violence: A meta-analysis on the effect of exposure and mental health outcomes of children and adolescents. Developmental Psychopathology 21, 227–259. Gapen, M., Cross, D., Ortigo, K., Graham, A., Johnson, E., Evces, M.,… Bradley, B. (2011).

Perceived neighborhood disorder, community cohesion, and PTSD symptoms among lowincome African Americans in an urban health setting. American Journal of Orthopsychiatry, 81(1), 31–37. doi: 10.1111/j.1939–0025.2010.01069.x Garbarino, J. (1996). The spiritual challenge of violent trauma. American Journal of Orthopsychiatry, 66(1), 162–163. doi: 10.1037/h0080167 Garbarino J., Dubrow, N., Kostelny, K. & Pardo, C. (1992). Children in danger: Coping with the effects of community violence. SanvFrancisco, CA: Jossey-Bass. Garrido, E. F., Culhane, S. E., Raviv, T. & Tasussig, H. N. (2010). Does community violence exposure predict trauma symptoms in a sample of maltreated youth in foster care? Ciolence and Victims, 25, 755–769. Gill, C., Weisburd, D., Telep, C., Vitter, Z., & Bennet, T. (2014). Community-oriented policing to reduce crime, disorder and fear and increase satisfaction and legitimacy among citizens: A systematic review. Journal of Experimental Criminology. 10, 399–428. doi: 10.1007/s11292–014–9210-y Gorman-Smith, D., Henry, D. B., & Tolan, P. H. (2004). Exposure to community violence and violence perpetration: The protective effects of family functioning. Journal of Clinical Child and Adolescent Psychology, 33(3), 439–449. doi: 10.1207/s15374424jccp3303 Gorman-Smith, D., & Tolan, P. H. (1998). The role of exposure to community violence and developmental problems among inner-city youth. Development and Psychopathology, 10(1), 101–116. doi: 10.1017/S0954579498001539 Gorman-Smith, D., Tolan, P. H., Sheidow, A. J., & Henry, D. (2002). Partner violence and street violence among urban adolescents: Do the same family factors relate? Journal of Research on Adolescence, 11(3), 273–295. doi: 10.1111/1532–7795.00013 Graczyk, P. A., & Tolan, P. H. (2005). Implementing effective youth violence prevention programs in community settings. In R. G. Steele & M. C. Roberts (Eds.), Handbook of mental health services for children, adolescents, and families (pp. 215–230). New York, NY: Kluwer Academic/Plenum. doi: 10.1007/0–387–23864–6_14 Groves, B., Zuckerman, B., Marans, S., & Cohen, D. (1993). Silent victims: Children who witness violence. Journal of the American Medical Association, 269, 262–264. doi: 10.1001/jama.269.2.262 Guerra, N. G., Huesmann, R. L., & Spindler, A. (2003). Community violence exposure, social cognition, and aggression among urban elementary school children. Child Development, 74, 1561–1576. Gunnar, M. R., & Vazquez, D. M. (2006). Stress neurobiology and developmental psychopathology. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology, Vol. 2: Developmental neuroscience (2nd ed., pp. 533–577). Hoboken, NJ: Wiley.

Hahn, R., Fuqua-Whitley, D., Wethington, H., Lowy, J., Crosby, A., Fullilove, M.,… & Dahlberg, L. (2007). Effectiveness of universal school-based programs to prevent violence and aggressive behavior: A systematic review. American Journal of Preventive Medicine, 33(Suppl.), S114–S129. Halliday-Boykins, C. A., & Graham, S. (2001). At both ends of the gun: Testing the relationship between community violence exposure and youth violent behavior. Journal of Abnormal Child Psychology, 29(5), 383–402 Hammack, P. L., Richards, M. H., Luo Z., Edlynn, E. S., & Roy, K. (2004). Social support factors as moderators of community violence exposure among inner-city African American young adolescents. Journal of Clinical Child and Adolescent Psychology, 33(3):450–462 Hanson, R. F., Self-Brown, S., Fricker-Elhai, A. E., Kilpatrick, D. G., Saunders, B. E., & Resnick, H. S. (2006). The relations between family environment and violence exposure among youth: findings from the national survey of adolescents. Child Maltreatment, 11, 3–15. Hardaway C. R., McLoyd V. C., & Wood, D. (2012). Exposure to violence and socioemotional adjustment in low-income youth: An examination of protective factors. American Journal of Community Psychology, 49(1–2), 112–126. doi: 10.1007/s10464–011–9440–3 Hawkins, J. D., Oesterle, S., Brown, E. C., Abbott, R. D., & Catalano, R. F. (2014). Youth problem behaviors 8 years after implementing the Communities That Care prevention system. A community-randomized trial. JAMA Pediatrics, 168(2), 122–129. doi: 10.1001/jamapediatrics.2013.4009 Henry, D. B., Gorman-Smith, D., Schoeny, M., & Tolan, P. H. (2014). “Neighborhood matters”: Assessment of neighborhood social processes. American Journal of Community Psychology, doi: 10.1007/s10464–014–9681-z Herrenkohl, T. I., Sousa, C., Tajima E. A., Herrenkohl, R. C., & Moylan, C. A. (2008). Intersection of child abuse and children's exposure to domestic violence. Trauma, Violence, & Abuse. 9, 84–99. doi: 10.1177/1524838008314797 Horney, J., Tolan, P. H., & Weisburd, D. (2012). Contextual influences. In R. Loeber & D. Farrington (Eds.), From juvenile delinquency to adult crime: Criminal careers, justice policy, and prevention (pp. 86–117). New York, NY: Oxford University Press. Horowitz, K., McKay, M., & Marshall, R. (2005). Community violence and urban families: experiences, effects, and directions for intervention. American Journal of Orthopsychiatry, 75(3), 56–68. Hudley, C., & Taylor, A. (2006). What is cultural competence and how can it be incorporated into preventive interventions? In N. G. Guerra & E. P. Smith (Eds.), Preventing youth violence in a multicultural society (pp. 249–269). Washington, DC: American Psychological Association. doi: 10.1037/11380–010

Huizinga, D., Haberstick, B. C., Smolen, A., Menard, S., Young, S. E., Corley, R. P.,… Hewitt, J. K. (2006). Childhood maltreatment, subsequent antisocial behavior, and the role of monoamine oxidase A genotype. Biological Psychiatry, 60(7), 677–683. doi: 10.1016/j.biopsych.2005.12.022 Jackman, M. R. (2002). Violence in social life. Annual Review of Sociology, 28, 387–415. doi: 10.1146/annurev.soc.28.110601.140936 James, B. (2011). Popular crime: Reflections on the celebration of violence. New York, NY: Scribner. Jaycox, L. H., Hickman, L. J., Schultz, D., Barnes-Proby, D., Setodji, J., Kofner, A., Harris, A., Acosta, J. D., & Francois, T. (2011). National evaluation of Safe Start Promising approaches: Assessing program outcomes. Santa Monica, CA: RAND Corporation, TR-991DOJ, 2011. http://www.rand.org/pubs/technical_reports/TR991.html Johnson, M. P. (1995). Patriarchal terrorism and common couple violence: Two forms of violence against women. Journal of Marriage and the Family, 57(2), 283–294. doi: 10.2307/353683 Jouriles, E. N., McDonald, R., Norwood, W. D., & Ezell, E. (2001). Issues and controversies in documenting the prevalence of children's exposure to domestic violence. In S. A. GrahamBermann & J. L. Edleson (Eds.), Domestic violence in the lives of children: The future of research, intervention, and social policy (pp. 12–34). Washington, DC: American Psychological Association. doi: 10.1037/10408–002 Kaufman J. M. (2009). Gendered responses to serious strain: The argument for a general strain theory of deviance. Justice Quarterly. 26, 410–444. Kaufman J., Yang, B. Z., Douglas-Palumberi, H., Houshyar, S., Lipschitz, D., Krystal, J. H., & Gelernter, J. (2004). Social support and serotonin transporter gene moderate depression in maltreated children. Proceedings of the National Academy Sciences, 101(49), 17316–17321. doi: 10.1073/pnas.0404376101 Kaynak, O., Lepore, S. J., & Kliewer, W. (2011) Social moderation of the relation between community violence exposure and depressive symptoms in an urban adolescent sample. Journal of Social and Clinical Psychology, 30(3), 250–269. Kendler, K. S., Kuhn, J. W., Vittum, J., Prescott, C. A., & Riley, B. (2005). The interaction of stressful life events and a serotonin transporter polymorphism in the prediction of episodes of major depression: A replication. Archives of General Psychiatry, 62(5), 529–535. doi: 10.1001/archpsyc.62.5.529 Kim-Cohen, J., Caspi, A., Taylor, A, Williams, B., Newcombe, R., Craig, I. W., & Moffitt, T. E. (2006). MAOA, maltreatment, and gene-environment interaction predicting children's mental health: New evidence and a meta-analysis. Molecular Psychiatry, 11(10), 903–913.

doi: 10.1038/sj.mp.4001851 Kitzmann, K. M., Gaylord, N. K., Holt, A. R., & Kenny, E. D. (2003). Child witnesses to domestic violence: a meta-analytic review. Journal of Consulting Clinical Psychology, 2, 339–352. Kliewer, W. (2006). Violence exposure and cortisol responses in urban youth. International Journal of Behavioral Medicine, 13(2), 109–120. doi: 10.1207/s15327558ijbm1302_2 Kliewer, W., Murrelle, L., Mejia, R., Torres, de Y., & Angold, A. (2001). Exposure to violence against a family member and internalizing symptoms in Colombian adolescents: The protective effects of family support. Journal of Consulting and Clinical Psychology, 69, 971– 82. Klorman, R., Cicchetti, D., Thatcher, J. E., & Ison, J. R. (2003). Acoustic startle in maltreated children. Journal of Abnormal Child Psychology, 31(4), 359–370. doi: 10.1023/A:1023835417070 Krenichyn, K., Saegert, S., & Evans, G. W. (2001). Parents as moderators of psychological and physiological correlates of inner-city children's exposure to violence. Applied Developmental Psychology, 22(6), 581–602. doi: 10.1016/S0193–3973(01)00095–8 Krug, E., Dahlberg, L., Mercy, J., Zwi, A., & Lozano, R. (2002). World report on violence and health. Geneva, Switzerland: World Health Organization. Lambert, S. F., Bettencourt, A. F., Bradshaw, C. P., & Ialongo, N. S. (2013). Early predictors of urban adolescents' community violence exposure. Journal of Aggression, Maltreatment & Trauma, 22, 26–44. doi: 10.1080/10926771.2013.743944 Lambert, S. F., Ialongo, N. S., Boyd, R. C., Cooley, M. R. (2005). Risk factors for community violence exposure in adolescence. American Journal of Community Psychology, 36 (1–2), 29–48. doi: 10.1007/s10464–005–6231–8 Lansford, J. E., Chang, L., Dodge, K. A., Malone, P. S., Oburu, P., Palmérus, K.,… Quinn, N. (2005). Physical discipline and children's adjustment: Cultural normativeness as a moderator. Child Development, 76(6), 1234–1246. doi: 10.1111/j.1467–8624.2005.00847.x Laub, J. H., and Lauritsen, J. L. (1998). The interdependence of school violence with neighborhood and family conditions. In D. S. Elliott, B. A. Hamburg, & K. R. Williams (Eds.), Violence in American schools (pp. 127–155). New York, NY: Cambridge University Press. Lauritsen, J., & Davis-Quinet, K. (1995). Repeat victimization among adolescents and young adults. Journal of Quantitative Criminology, 11(2), 143–166. Leenarts, L. E., Diehle, J., Doreleijers, T. A., Jansma, E. P., & Lindauer, R. J. (2013). Evidence-based treatments for children with trauma-related psychopathology as a result of childhood maltreatment: A systematic review. European Child & Adolescent Psychiatry. 22,

269–83. doi: 10.1007/s00787–012–0367–5 Leventhal, T., & Brooks-Gunn, J. (2000). The neighborhoods they live in: The effects of neighborhood residence on child and adolescent outcomes. Psychological Bulletin, 126(2), 309–337. doi: 10.1037/0033–2909.126.2.309 Lieberman, A. Chu, A., Van Horn, P., & Harris, W. (2011). Trauma in early childhood: Empirical evidence and clinical Implications. Development and Psychopathology, 23, 397– 410. Lieberman, A., & DeMartino, R. (2006). Interventions for children exposed to violence. Johnson & Johnson Institute. New Brunswcik, NJ. Loseke, D. R., Gelles, R. J., & Cavanaugh, M. M. (Eds.). (2005). Current controversies on family violence. (2nd ed.). Thousand Oaks, CA: Sage. Low, S., & Espelage, D. (2014). Conduits from community violence exposure to peer aggression and victimization: Contributions of parental monitoring, impulsivity, and deviancy. Journal of Counseling Psychology. Advance online publication. 10.1037/a0035207 Lupien S. J., King S., Meaney M., & McEwen B. (2001). Can poverty get under your skin? Basal cortisol levels and cognitive function in children from low and high socioeconomic status. Development and Psychopathology,13, 653–676. Lynch, M. (2003). Consequences of children's exposure to community violence. Clinical Child and Family Psychology Review, 6, 265–274. Lynch, M., & Cicchetti, D. (1998). An ecological-transactional analysis of children and contexts: The longitudinal interplay among child maltreatment, community violence, and children's symptomatology. Development and Psychopathology, 10, 235–257. Lynch, M., & Cicchetti, D. (2002). Links between community violence and the family system: Evidence from children's feelings of relatedness and perceptions of parent behavior. Family Process, 41(3), 519–532. MacMillan, R & Hagan, J. (2004). Violence in the transition to adulthood: Adolescent victimization, education and socioeconomic attainment in later life. Journal of Research on Adolescence, 14, 127–158. DOI: 10.1111/j.1532-7795.2004.01402001.x Maikovich, A. K., Jaffee, S. R., Odgers, C. L., & Gallop, R. (2008). Effects of family violence on psychopathology symptoms in children previously exposed to maltreatment. Child Development, 79(5), 1498–1512. doi: 10.1111/j.1467–8624.2008.01201.x. Makarios, M. D., & Pratt, T. C. (2012). The effectiveness of policies and programs that attempt to reduce firearm violence: A meta-analysis. Crime & Delinquency 58(2), 222–244. Malik, N. M. (2008). Exposure to domestic and community violence in a non-risk sample: Associations with child functioning. Journal of Interpersonal Violence, 23(4), 490–504. doi:

10.1177/0886260507312945. Malley-Morrison, K., & Hines, D. (2004). Family violence in a cultural perspective: Defining, understanding, and combating abuse. Thousand Oaks, CA: Sage. Marans, S., & Berkman, M., (2006). Police-mental health collaboration on behalf of children exposed to violence: The Child Development-Community Policing Model. A. Lightburn & P. Sessions (Eds.), Handbook of community-based clinical practice (pp. 426–440). New York, NY: Oxford University Press. Margolin, G., & Gordis, E. B. (2000). The effects of family and community violence on children. Annual Review of Psychology, 51, 445–479. doi: 10.1146/annurev.psych.51.1.445 Margolin, G., & Gordis, E. B. (2004). Children's exposure to violence in the family and community. Current Directions in Psychological Science, 13, 152–155. doi: 10.1111/j.0963– 7214.2004.00296.x Martinez, P., & Richters, J. (1993). The NIMH Community Violence Project: II. Children's distress symptoms associated with violence exposure. Psychiatry, 56, 22–35. McCabe, K. M., Lucchini, S. E., Hough, R. L., Yeh, M., & Hazen, A. (2005). The relation between violence exposure and conduct problems among adolescents: A prospective study. American Journal of Orthopsychiatry, 75(4), 575–584. doi: 10.1037/0002–9432.75.4.575 McCart, M., Smith, D. W., Saunders, B. E., Kilpatrick, D. G., Resnick, H., & Ruggiero, K. J. (2007). Do urban adolescents become desensitized to community violence? Data from a national survey. American Journal of Orthopsychiatry, 77(3), 434–442. doi: 10.1037/0002– 9432.77.3.434 McGee, Z. T. & Baker, S. R. (2003). Impact of violence on problem behavior among adolescents. Journal of Contemporary Criminal Justice, 18, 74–93. doi: 10.1177/1043986202018001006 Mead, H. K., Beauchaine, T. P., & Shannon, K. E. (2010). Neurobiological adaptations to violence across development. Development and Psychopathology, 22(1), 1–22. doi: 10.1017/S0954579409990228 Miller, L., Wasserman, G., Neugebauer, R., Gorman-Smith, D., & Kamboukos, D. (1999). Witnessed community violence and anti-social behavior in high-risk, urban boys. Journal of Clinical Child Psychology, 28(1), 2–11. doi: 10.1207/s15374424jccp2801_1 Morales, J. R., & Guerra, N. G., (2006). Effects of multiple context and cumulative stress on urban children's adjustment in elementary school. Child Development, 77(44), 907–923. Morgan, C. A., Grillon, C., Lubin, H., & Southwick, S. M. (1997). Startle reflex abnormalities in women with sexual assault-related posttraumatic stress disorder. American Journal of Psychiatry, 154(8), 1076–1080.

Morgan, C. A., Grillon, C., Southwick, S. M., Davis, M., & Charney, D. S. (1996). Exaggerated acoustic startle reflex in Gulf War veterans with posttraumatic stress disorder. American Journal of Psychiatry. 153, 64–68. Mrug, S., Loosier, P. S., & Windle, M. (2008). Violence exposure across multiple contexts: Individual and joint effects on adjustment. American Journal of Orthopsychiatry, 78(1), 70– 84. doi: 10.1037/0002–9432.78.1.70 Multisite Violence Prevention Project. (2009). The ecological effects of universal and selective violence prevention programs for middle school students: A randomized trial. Journal of Consulting and Clinical Psychology, 77(3), 526–542. doi: 10.1037/a0014395 Multisite Violence Prevention Project. (2014). The impact of universal versus selective approaches to prevention on high-risk, socially-influential middle school students. Journal of Research in Adolescence, 24, 364–382. Nader K., Pynoos R., Fairbanks L., Frederick, C. (1990). Children's PTSD reactions one year after a sniper attack at their school. American Journal of Psychiatry, 147(11), 1526–1530. National Research Council of the National Academies. (2004). Firearms and violence: A critical review. Washington, DC: National Academies Press. National Research Council of the National Academies. (2013). Priorities for research to reduce the threat of firearm related violence. Washington, DC: National Academies Press. Office of the Surgeon General. (2001). Youth violence: A Report of the Surgeon General. Rockville, MD: Author. Available from: http://www.ncbi.nlm.nih.gov/books/ Overstreet, S., & Braun, S. (2000). Exposure to community violence and post-traumatic stress symptoms: Mediating factors. American Journal of Orthopsychiatry, 70, 263–271. Overstreet, S., & Mazza, J. (2003). An ecological-transactional understanding of community violence: Theoretical perspectives. School Psychology Quarterly, 18(1), 66–87. Paolucci, E. O., Genuis, M. L., & Violato, C. (2001). A meta-analysis of the published research on the effects of child sexual abuse. Journal of Psychology, 135(1), 17–36. Park-Higgerson, H.-K., Perumean-Chaney, S. E., Bartolucci, A. A., Grimley, D. M., & Singh, K. P. (2008). The evaluation of school-based violence prevention programs: A meta-analysis. Journal of School Health, 78, 465–479. Peckins, M. K., Dockray, S., Eckenrode, J. L., Heaton, J., & Susman, E. J. (2012). The longitudinal impact of exposure to violence on cortisol reactivity in adolescents. Journal of Adolescent Health, 51(4), 366–372. doi: 10.1016/j.jadohealth.2012.01.005 Perkins, S., & Graham-Bermann, S. (2012). Violence exposure and the development of schoolrelated functioning: Mental health, neurocognition, and learning. Aggression and Violent Behavior, 17(1), 89–98.

Perry, B. D. (2008). Child maltreatment: The role of abuse and neglect in developmental psychopathology. In T. P. Beauchaine & S. P. Hinshaw (Eds.), Child and adolescent psychopathology (pp. 93–128). Hoboken, NJ: Wiley. Phillips, R. G., & LeDoux, J. E. (1992). Differential contribution of amygdala and hippocampus to cued and contextual fear conditioning. Behavioral Neuroscience, 106(2), 274–285. doi: 10.1037/0735–7044.106.2.274 Pinchevsky, G. M., Wright, E. M., & Fagan, A. A. (2013). Gender differences in the effects of exposure to violence on adolescent substance use. Violence and Victims, 28(1), 122–144. Pine, D. S., Coplan, J. D., Wasserman, G. A., Miller, L. S., Fried, J. E., Davies, M,… Parsons, B. (1997). Neuroendocrine response to fenfluramine challenge in boys: Associations with aggressive behavior and adverse rearing. Archives of General Psychiatry, 54(9), 839–846. doi: 10.1001/archpsyc.1997.01830210083010 Plotsky, P. M., & Meaney, M. J. (1993). Early, postnatal experience alters hypothalamic corticotropin-releasing factor (CRF) mRNA, median eminence CRF content and stress-induced release in adult rats. Molecular Brain Research, 18(3), 195–200. Plybon, L. E., & Kliewer, W. (2001). Neighborhood types and externalizing behavior in urban school-age children: Tests of direct, mediated, and moderated effects. Journal of Child and Family Studies, 10, 419–437. Pope, A. W., & Bierman, K. L. (1999). Predicting adolescent peer problems and antisocial activities: The relative roles of aggression and dysregulation. Developmental Psychology, 35, 335–346. Preski, S., & Shelton, D. (2001). The role of contextual, child and parent factors in predicting criminal outcomes in adolescence. Issues in Mental Health Nursing, 22(2), 197–205. Raine, A., Moffitt, T. E., Caspi, A., Loeber, R., Stouthamer-Loeber, M., & Lynam, D. (2005). Neurocognitive impairments in boys on the life-course persistent antisocial path. Journal of Abnormal Psychology, 114(1), 38–49. doi: 10.1037/0021–843X.114.1.38 Ratner, H. H., Chiodo, L., Covington, C., Sokol, R. J., Ager, J., & Delaney-Black, V. (2006). Violence exposure, IQ, academic performance, and children's perception of safety: evidence of protective effects. Merrill-Palmer Quarterly, 52(2), 264–287. doi: 10.1353/mpq.2006.0017 Rumm, P., Cummings, P., Krauss, M., Bell, M., & Rivara, F. (2000). Identified spouse abuse as a risk factor for child abuse. Child Abuse and Neglect, 24, 1375–1381. Salzinger, S., Feldman, R., Stockhammer, T., & Hood, J. (2002). An ecological framework for understanding risk for exposure to community violence and the effects of exposure on children and adolescents. Aggression and Violent Behavior, 7(5), 423–451. Sampson, R. J. (1997). Collective regulation of adolescent misbehavior: Validation results

from eighty Chicago neighborhoods. Journal of Adolescent Research, 12(2), 227–244. doi: 10.1177/0743554897122005 Sampson, R. J., Raudenbush, S. W., & Earls, F. (1997). Neighborhoods and violent crime: A multilevel study of collective efficacy. Science, 277, 918–924. Schwab-Stone, M., Ayers, T., Kasprow, W., Voyce, C., Barone, C., Shriver, T., & Weissberg, R. (1995). No Safe Haven: A study of violence exposure in an urban community. Journal of the American Academy of Child and Adolescent Psychiatry, 10, 1343–1352. Schwab-Stone, M., Chen, C., Greenberger, E., Silver, D., Lichtman, J., & Voyce, C. (1999). No safe haven II: The effects of violence exposure on urban youth. Journal of the American Academy of Child and Adolescent Psychiatry, 38(4), 359–367. Schwartz, D., & Gorman, A. H. (2003). Community violence exposure and children's academic functioning. Journal of Educational Psychology, 95(1), 163–173. doi: 10.1037/0022– 0663.95.1.163 Schwartz, D., & Proctor, L. (2000). Community violence exposure and children's social adjustment in the school peer group: The mediating roles of emotion regulation and social cognition. Journal of Consulting and Clinical Psychology, 68(4), 670–668. doi: 10.1037/0022–006X.68.4.670 Shalev, I., Moffitt, T. E., Sugden, K., Williams, B., Houts, R. M., Danese, A.,… Caspi, A. (2013). Exposure to violence during childhood is associated with telomere erosion from 5 to 10 years of age: A longitudinal study. Molecular Psychiatry, 8, 576–81. doi: 10.1038/mp.2012.32 Sharkey, P. T., Tirado-Strayer, N., Papachristos, A. V., and Raver, C. C. (2012). The effect of local violence on children's attention and impulse control. American Journal of Public Health, 102(12), 2287–2293. doi: 10.2105/AJPH.2012.300789. Epub 2012 Oct 18. Sheidow, A., Gorman-Smith, D., Tolan, P. H. and Henry, D. (2001). Family and community characteristics: risk factors for violence exposure in inner-city youth. Journal of Community Psychology, 29, 345–60. Shields, N., Nadasen, K. & Pierce, L. (2009). A comparison of the effects of witnessing community violence and direct victimization among children in Cape Town, South Africa. Journal of Interpersonal Violence, 24(7), 1192–1208. doi: 10.1177/0886260508322184 Spano, R., Rivera, C., & Bolland, J. M. (2010). Are chronic exposure to violence and chronic violent behavior closely related developmental processes during adolescence? Criminal Justice and Behavior, 37(10), 1160–1179. Stein, B. D., Jaycox, L. H., Kataoka, S., Rhodes, H. J., & Vestal, K. D. (2003). Prevalence of child and adolescent exposure to community violence. Clinical Child and Family Psychology Review, 6(4), 247–264. doi: 10.1023/B:CCFP.0000006292.61072.d2

Stein, B. D., Jaycox, L. H., Kataoka, S., Wong, M., Tu, W., Elliott, M. N., & Fink, A. (2003). A mental health intervention for school children exposed to violence: A randomized controlled trial. Journal of the American Medical Association, 290, 603–611 Swaner, R., & Kohn, J. (2011, November). The U.S. Attorney General's Defending Childhood Initiative: formative evaluation of the Phase 1 Demonstration Program. Final report to the National Institute of Justice, Grant Number 2010-IJ-CX-0015, NCJ 236563. Thompson, T., Jr. & Massat, C. R. (2005). Experiences of violence, post-traumatic stress, academic achievement and behavior problems of urban African American children. Child & Adolescent Social Work Journal, 22(5–6), 367–393. doi: 10.1007/s10560–005–0018–5 Thornberry, T. P. (1998). Membership in youth gangs and involvement in serious and violent offending. In R. Loeber & D. P. Farrington (Eds.), Serious and violent juvenile offenders: Risk factors and successful interventions (pp. 147–166). Thousand Oaks, CA: Sage. Tolan, P. H. (2001a). Emerging themes and challenges in understanding youth violence involvement. Journal of Clinical Child Psychology, 30(2), 233–239. doi: 10.1207/S15374424JCCP3002_10 Tolan, P. H. (2001b). Youth violence and its prevention in the United States: An overview of current knowledge. Journal of Inquiry Control and Safety Promotion, 8(1), 1–12. doi: 10.1076/ icsp.8.1.1.3374 Tolan, P. H. (2007). Understanding violence. In I. D. Waldman (Ed.), The Cambridge handbook of violent behavior and aggression (pp. 5–18). New York, NY: Cambridge University Press. Tolan, P. H., Dodge, K., & Rutter, M. (2013). The multiple pathways of family influence on disruptive behavior problems. In P. Tolan & B. Leventhal (Eds.), Advances in development and psychopathology. Brain Research Foundation Symposium series, Volume I: Disruptive behavior problems (pp. 161–192). New York, NY: Springer. Tolan, P. H., & Gorman-Smith, D. (1997). Families and the development of urban children. In J. Herbert, O. R. Wallberg, & R. P. Weissberg (Eds.), Children and youth: Interdisciplinary perspectives (pp. 67–91). Thousand Oaks, CA: Sage Publications, Inc. Tolan, P. H., & Gorman-Smith, D. (2002). What violence prevention research can tell us about developmental psychopathology. Development and Psychopathology, 14(4), 713–729. doi: 10.1017/S0954579402004042 Tolan, P. H., Gorman-Smith, D., & Henry, D. B. (2003). The developmental ecology of urban males' youth violence. Developmental Psychology, 39(2), 274–291. doi: 10.1037/0012– 1649.39.2.274 Tolan, P. H., Gorman-Smith, D., & Henry, D. (2004). Supporting families in a high-risk setting: Proximal effects of the SAFE Children preventive intervention. Journal of Consulting and

Clinical Psychology, 72(5), 855–869. doi: 10.1037/0022–006X.72.5.855. Tolan, P. H., Gorman-Smith, D., & Henry, D. (2006). Family violence. Annual Review of Psychology, 57, 557–583. doi: 10.1146/annurev.psych.57.102904.190110 Tolan, P. H., & Grant, K. (2009). How social and cultural context shapes the development of coping: Youth in the inner-city as an example. In E. A. Skinner & M. J. Zimmer-Gembeck (Eds.), New directions for child and adolescent development (pp. 61–74). doi: 10.1002/cd.243 Tolan, P. H., & Guerra, N. (1998). Societal causes of violence against children. In P. K. Trickett & C. J. Schellenbach (Eds.), Violence against children in the family and the community (pp. 195–209). Washington, DC: American Psychological Association. doi: 10.1037/10292–007 Tolan, P. H., Guerra, N. G., & Kendall, P. C. (1995). A developmental-ecological perspective on antisocial behavior in children and adolescents: Toward a unified risk and intervention framework. Journal of Consulting and Clinical Psychology, 63(4), 579–584. doi: 10.1037/0022–006X.63.4.579 Tolan, P. H., Guerra, N. G., & Montaini-Klovdahl, L. (1997). Staying out of harm's way: Coping and the development of inner-city children. In S. A. Wolchik & I. N. Sandler (Eds.), Handbook of children's coping: Linking theory and intervention (pp. 453–479). New York, NY: Plenum Press. Tolan, P. H., & Morris, K. (2005). The inner-city of the United States and risk for substance use among youth. In I. S. Obot & S. Saxena (Eds.), Substance use among young people in urban environments (pp. 77–102). Geneva, Switzerland: World Health Organization, Department of Mental Health and Substance Abuse; WHO Centre for Health Development. Tolan, P. H., Sherrod, L. R., Gorman-Smith, D., & Henry, D. B. (2004). Building protection, support, and opportunity for inner-city children and youth and their families. In A. L. Solarz (Ed.), Investing in children, youth, families, and communities: Strengths-based research and policy (pp. 193–211). Washington, DC: American Psychological Association. doi: 10.1037/10660–011 Tremblay, R. E., Nagin, D. S., Seguin, J. R., Zoccolillo, M., Zelazo, P. D., Boivin, M.,… Japel, C. (2005). Physical aggression during early childhood: Trajectories and predictors. Canadian Child and Adolescent Psychiatry Review, 14(1), 3–9. Trentacosta C. J., Hyde L. W., Shaw, D. S., & Cheong, J. (2009). Adolescent dispositions for antisocial behavior in context: The roles of neighborhood dangerousness and parental knowledge. Journal of Abnormal Psychology. 118, 564–575. Trickett, P. K., & McBride-Chang, C. (1995). The developmental impact of different forms of child abuse and neglect. Developmental Review, 15(3), 311–337. doi:

10.1006/drev.1995.1012 Ttofti, M., & Farrington, D. (2010). School-based programs to reduce bullying and victimization. Campbell Systematic Reviews, 6. doi: 10.4073/csr.2009.6 Twardosz, S., & Lutzker, J. R. (2010). Child maltreatment and the developing brain: A review of neuroscience perspectives. Aggression and Violent Behavior, 15(1), 59–68. doi: 10.1016/j.avb.2009.08.003 Walters, R. H., & Parke, R. D. (1964). Influence of response consequences to a social model on resistance to deviation. Journal of Experimental Child Psychology, 1(3), 269–280. doi: 10.1016/0022–0965(64)90042–6 Watts, R. J., Abdul-Adil, J. K., & Pratt, T. (2002). Enhancing Critical Consciousness in Young African American Men: A Psychoeducational Approach. Psychology of Men & Masculinity, 3(1), 41–50. White, K. S., Bruce, S. E., Farrell, A. D., & Kliewer, W. (1998). Impact of exposure to community violence on anxiety: A longitudinal study of family social support as a protective factor for urban adolescents. Journal of Child and Family Studies, 7, 187–203. Widom, C. S., & Brzustowicz, L. M. (2006). MAOA and the “cycle of violence”: Childhood abuse and neglect, MAOA genotype, and risk for violent and antisocial behavior. Biological Psychiatry, 60(7), 684–689. doi: 10.1016/j.biopsych.2006.03.039 Wilson, D., Kliewer, W., Teasley, N., Plybon, L., & Sica, D. (2002). Violence exposure, catecholamine excretion, and blood pressure non-dipping status in African American male versus female adolescents. Psychosomatic Medicine, 64, 906–915. Wilson, J. (1987). The truly disadvantaged: The inner city, the underclass, and public policy. Chicago, IL: University of Chicago. doi: 10.1097/01.PSY.0000024234.11538.D3 Wilson, M. (1988). Marital conflict and homicide in evolutionary perspective. In R. W. Bell N. J. Bell (Ed.), (pp. 45–62). Lubbock, TX: Texas Tech University Press. World Health Organization Summit on Violence (2002). World report on violence and health. Geneva, Switzerland: World Health Organization. Wycoff, M., & Skogan, W. (1993). Community policing in Madison: Quality from the inside out. An evaluation of implementation and impact. Research report. Washington, DC: National Institute of Justice, U.S. Department of Justice. Zielinski, D. S. (2009). Child maltreatment and adult socioeconomic well-being. Child Abuse and Neglect, 33, 666–678. doi: 10.1016/j.chiabu.2009.09.001. Zinzow, H. M., Ruggiero, K. J., Hanson, R. F., Smith, D. W., Saunders, B. E., & Kilpatrick, D. G. (2009). Witnessed community and parental violence in relation to substance use and delinquency in a national sample of adolescents. Journal of Traumatic Stress, 22, 525–533.

doi: 10.1002/jts.20469

Chapter 3 Social Support and Developmental Psychopathology Ross A. Thompson and Rebecca Goodvin We are grateful for very helpful comments from Kristina Gelardi on an earlier version of this chapter. DEFINING SOCIAL SUPPORT Functions of Social Support for Adults and Children Social Monitoring as Social Support What Social Support Is and Is Not SOCIAL SUPPORT AND STRESS Social Buffering of Biological Stress Reactions Stress and Access to Social Support SOCIAL SUPPORT IN RELATIONSHIPS AND SOCIAL NETWORKS The “Active Ingredients” of Supportive Relationships Relationships and the Social Networks of Parents and Children Development and Functioning of Children's Social Networks Online Social Support GIVING AND RECEIVING SOCIAL SUPPORT IN A CULTURAL CONTEXT Recipient and Helper Reactions to Assistance Cultural Considerations in Giving and Receiving Social Support INTERIM CONCLUSIONS SOCIAL SUPPORT AND DEVELOPMENTAL PSYCHOPATHOLOGY Social Support, Social Isolation, and the Origins of Developmental Psychopathology Social Support and the Maintenance of Developmental Psychopathology Social Support and the Prevention and Treatment of Developmental Psychopathology CONCLUSIONS AND FUTURE DIRECTIONS REFERENCES We live in an environment of relationships. Those relationships are important for many reasons, and one of the most important is the psychological support they provide. Family members can be counted on for understanding and advice, help with child care, and financial

assistance. Friends are sources of encouragement and counseling, for sharing information and experience, and as referral sources. Coworkers can provide professional advice and guidance about workplace concerns. Neighbors are valuable sources of information and tangible aid, sometimes in the exchange of favors or resources. Marital and romantic partners support selfesteem and emotional well-being. Although not all relationships function optimally, and thus these forms of assistance may be provisional, the value of these relationships rests on their offering support in the context of the emotional ties that define them. This is especially true when individuals need help because they are under stress. Formal helping relationships are often intended to emulate these forms of informal support. Counselors, religious advisors, social workers, and other helping professionals attend to the psychological needs of their advisees as well as their material or spiritual condition. Teachers, professional coaches, and personal advisors likewise provide encouragement and support in the context of promoting skill development and knowledge acquisition. Social support is a critical function of many forms of psychological therapy, whether support is received from an individual mental health professional, a therapy or support group, or a special preschool, classroom, or adult education program designed for troubled individuals. Although individuals benefit from the support received in formal as well as informal helping relationships on any occasion, when they face significant stress or psychological turmoil, the psychological support that formal helping relationships afford can be critical to their coping and recovery. Children depend even more significantly than adults on the support provided by their environment of relationships. This is because central developmental achievements—emerging sense of self, social and emotional competence, cognitive and language skills, even physical health and development—are constructed in the context of social interactions in which the supportive assistance of relational partners is critical. In addition, because children (especially when they are young) are less capable of creating alternative sources of support for themselves, they rely more significantly on support from people in the family, school, and neighborhood social networks that are created for them by their parents. This is one reason why social support to children differs from social support for adults: It is mediated to a significant extent by parents and the opportunities they provide. Parents are also significant to children's support because they are key providers of the varieties of emotional, informational, material, and other resources that children need. They are, as we subsequently describe, support generalists. Where children are concerned, therefore, it is essential to think multigenerationally about relationships and social support. Understanding the value and functions of social support in everyday life provides a basis for understanding its role in developmental psychopathology. As described in this chapter, social support is relevant to developmental psychopathology in at least three ways. First, social support or its absence is important to the origins of clinical problems for children and their families. Children who encounter adversity without the buffering assistance of supportive relationships are at higher risk of developing psychopathology. Risk is especially heightened when family relationships become sources of adversity rather than of security because of children's reliance on the support of family members. For this reason, many preventive programs, such as home visitation, enlist social support to improve family conditions for

children, using the assistance of formal helpers to strengthen informal support to children. Second, social support or its absence is important to maintaining clinical symptomatology for children in troubled families. In this sense, family ecologies can become enabling conditions for affective psychopathology because of the absence of the kinds of support from family members that might help remediate children's anxiety, conduct problems, or other difficulties, and these family conditions instead exacerbate these problems. Third, social support is important to the treatment of clinical problems of children. Indeed, it is arguable that most forms of therapeutic assistance entail some form of social support, so the important questions concern the sources of support, the avenues by which it is provided, and most important the clinical symptomatology it is expected to address. Several models of therapeutic intervention entailing different kinds of supportive assistance are profiled later in this chapter. Taken together, the value of social support seems obvious and obviously beneficial in everyday experience, and thus of clear relevance to the prevention and treatment of clinical problems in children. Such a view led to what Shumaker and Brownell (1984) described as an “era of unrestrained enthusiasm that has dominated social support research and interventions for more than 10 years.” What led, then, to the close of this enthusiastic period? Some considerations are listed next. Social support may address some of the emotional needs of its recipients without addressing the root problems that contribute to dysfunctional behavior. This problem is evocatively illustrated by Jill Korbin's (1989, 1991, 1995) interviews with mothers convicted of fatal child abuse. These mothers were surrounded by family, friends, and neighbors who were often painfully aware of the bruises, neglect, and other harms the mothers inflicted on their offspring. But in their efforts to be emotionally supportive, they failed to challenge abusive practices and instead overlooked other signs of parental dysfunction, minimized the seriousness of abuse, and offered reassurance about the mothers' good intentions while providing noncritical emotional affirmation. In doing so, of course, they contributed little to curbing abusive practices or protecting the children. As Korbin (1991, p. 23) noted, “A high level of perceived support sustained, probably unintentionally, these women in their pattern of abusive behavior.” It is not difficult to see similar patterns of emotional support enabling other forms of family pathology that victimize children, such as family practices that excuse rather than challenge parental substance abuse or domestic violence by offering reassurance without addressing the causes of clinical symptomatology. Social support may have attendant negative consequences that ultimately undermine sustaining support to needy recipients. Home visitors who seek to provide emotional support, material assistance, and information to at-risk parents are often surprised, for example, to discover that parents are absent during scheduled visits and resistant to the assistance they receive, and that they respond to the home visitor not with gratitude but instead with resentment and ill will. The home visitors do not understand that receiving one-way assistance, even (perhaps especially) from someone motivated by beneficent motives, often generates feelings of humiliation, failure, stigmatization, and vulnerability in recipients (J. D. Fisher, Nadler, & Witcher-Alagna, 1982; Schumaker & Brownell, 1984).

One implication is that how social support is provided and the nature of the relationship in which it is offered are crucial to recipient reactions to assistance and the maintenance of a helping association. Without considering recipient reactions to assistance, helping relationships often break down. Social support may exhaust the resources of informal and formal helpers, and the characteristics of those most in need of social support are especially likely to drain helping relationships. Friendships become stretched by the emotional neediness of someone who is in chronic distress, and individuals may begin to distance themselves from those who are depressed because of the negative affect, demandingness, and even hostility that can become part of their relationship. Troubled families with neglected children may evoke compassion in neighbors but also aversion because of parents' disorganization, social incompetence, and withdrawal (Gaudin & Polansky, 1986; Polansky & Gaudin, 1983). Professional clinicians also experience the drain of providing emotional assistance to their clients. The maintenance of supportive relationships, whether informal or formal in nature, requires attention to the needs of helpers as well as recipients. The current era of social support research and interventions has learned these lessons, and as a consequence, there is greater attention to tailoring supportive interventions to the needs of specific recipient populations, devoting attention to the contexts in which assistance is offered, considering the characteristics of helpers and those they aid, and creating networks in which social support is most likely to be maintained (Thompson, 2015a). In addition, social support in the context of developmental psychopathology increasingly relies on understanding children's social networks and the role of parents as organizers, gatekeepers, and mediators of children's access to helpers. Research in developmental relational science provides insight into the influence of relationships to promote or undermine children's well-being. Recognizing that social support is easy to conceptualize but difficult to implement in the lives of troubled children and their families, the field of social support research and intervention is less enthusiastic, more cautious, but potentially more helpful in addressing the clinical needs of children and families than it was several decades earlier (Thompson, 2015a). The purpose of this chapter is to consider the complex and multifaceted association between social support and developmental psychopathology. Our goal is to draw on diverse research literatures to illustrate how social support is relevant to the etiology, maintenance, and treatment of clinical disorders in children and families in the context of a developmental appraisal of children's social support networks. Our goal is also to use the lessons of treatment interventions and basic research to outline suggestions for how social support interventions can have the most positive effects for children and families as well as to outline directions for future research. The chapter opens with a discussion of how to define and conceptualize social support, considering both what it is and what it is not. In doing so, the multiple functions of social support for adults and children are discussed, with attention to similarities and differences in the experience of social support for children and adults. Regardless of age, however, social support is believed to contribute to psychological well-being in two ways: by preventing

stresses from occurring, and by buffering the effects of stress when it occurs. The next section of the chapter discusses multiple ways by which support has these consequences, including studies of the social buffering of stress neurobiology (Thompson, 2014a). Not surprisingly, the association between social support and stress is complex and bidirectional, and these bidirectional associations also are considered. Recognizing that social support is provided most often in the context of relationships and that relationships are multifaceted, the next section considers the characteristics of relationships that afford support, as well as how relationships sometimes can undermine support between relational partners. Also discussed is how obtaining support in a relational context compares for children and adults and how children's access to support is mediated by parents and family conditions. Developmental changes in children as relational partners are also discussed, especially how these changes increasingly enable children to construct and maintain their own social networks. The nature of social support in the context of online relationships through social media and other technological avenues is considered as well. Then this relational perspective is extended to consider cultural processes related to giving and receiving social support. A series of interim conclusions about social support then follow. In the next section, we discuss most directly the relevance of social support to risk for affective psychopathology in children in relation to the origins, maintenance, and treatment of clinical disorders. We use examples from well-studied areas of clinical psychopathology or atrisk family conditions, rather than focusing directly on a few diagnostic categories, in the hope that this will yield more general lessons for those who seek to enlist social support interventions to aid at-risk children and families. We also offer examples of promising intervention efforts as exemplary models. These lessons learned provide the transition to the final section, which offers directions for future research and intervention.

Defining Social Support To most adults, “social support” consists of the emotional affirmation received from close relational partners, particularly during periods of distress or turmoil. Most researchers believe that although emotional affirmation is important, social support is actually a more multifaceted phenomenon (Gottlieb & Bergen, 2010). The next definition reflects this fact: Social support consists of social relationships that provide (or can potentially provide) material and interpersonal resources that are of value to the recipient, such as counseling, access to information and services, sharing of tasks and responsibilities, and skill acquisition. (Thompson, 1995, p. 43) Unpacking this definition provides a more complex portrayal of the functions of social support and better understanding of how social support to children compares with support to adults.

Functions of Social Support for Adults and Children First, social support offers emotional encouragement, including affirmation, understanding, and empathy, the sense that others share one's troubles and are reliably “on your side” (what

Gottlieb, 1985, calls “milieu reliability”). For adults, this emotional function contributes to a sense of perceived support from social network members, especially those in close relationships, that is an important component of how support buffers stress (Barrera, 1986; Cohen & Wills, 1985). The emotional understanding and companionship entailed in this support function is important also to children. From infancy, children rely on the solicitude and emotional support of their caregivers and develop secure attachments to those adults based on the sensitivity and reliability of the adult's assistance (Cassidy, 2008; Thompson, 2006). The security in a caregiver's support and responsiveness is important whether children's relationships are with adults within or outside the family, and it contributes to children's social and emotional competence and self-confidence (Thompson, 2008). Likewise, children's friendships incorporate emotional understanding and acceptance, especially as children mature and as intimate self-disclosure and emotional support become increasingly important to friendship quality (Gottman & Parker, 1987). Second, social support provides counseling, advice, and guidance. For adults, counseling may be sought concerning marital or child care concerns, workplace issues, personal problems, or other issues. Children require advice and guidance for a much broader range of individual and developmental issues because caregivers have much greater knowledge and experience about major life issues, and peers are sources of relevant social understanding and perspective. This function of social support is thus, like emotional support, essential to healthy child development (Thompson, 2014b). The counseling and guidance function of social support may be combined with emotional encouragement, but whether this occurs or not depends on the nature of the relationship between the recipient and the support agent. Adults may seek guidance from a workplace colleague or neighbor about a problem without expecting emotional support, for example, but they may obtain advice with the expectation of encouragement from a spouse. Likewise, children may obtain helpful advice from a teacher, peer, or parent in the context of different expectations for emotional support. Indeed, social discriminations concerning expectations for the kind of support that might be solicited from different members of one's social network become part of social-cognitive growth in middle childhood and adolescence. During this period, for example, peers become significant sources of advice on matters of style (hair, dress, music) and activities, whereas parents remain important for guidance on questions of moral values and future goals (Coleman, 2011). In this respect, different network associates are preferred for different kinds of advice, which may or may not be accompanied by emotional affirmation. Third, social support enables access to information, services, and material resources and assistance. In this way, support agents provide tangible aid or act as brokers with others who can do so. For adults, this may occur when neighbors and friends are consulted for information about schools, community agencies, or child care settings or when extended family lend money or provide needed transportation. These activities often overlap with another function of social support, sharing of tasks and responsibilities, in which people assist the recipient in these ways or provide access to others who can do so. For children, these functions of social support are more extensive because children depend on adults within and outside the home for

information, services, and the necessities of life. This assistance can occur in the informal settings of family or peer group, or through the formal services of school or community agencies. Importantly, however, access to these formal services depends, at least until children are older, on the permission and assistance of parents. Fourth, social support can contribute to skill acquisition. Skill acquisition occurs for adults when coworkers assist with job-related skills, neighbors assist in learning about gardening or car repair, and friends or extended family aid in personal skills related to household management, parenting, or financial planning. Skill acquisition is a central feature of social support for children, whether considering parents, peers, or people outside the family (such as teachers or members of a faith community) as support agents. As in other support functions, the contributions of others to children's skill acquisition is more extensive than for adults because many of the abilities to which parents and peers contribute are developmentally foundational. Social skills, literacy, navigating familiar routes, and consumer practices are among the multitude of skills that are acquired in the company of social partners. As with sources of advice and guidance, children acquire different skills from different people in their support networks. Peer relationships are, for example, rich contexts for the acquisition of unique social skills that are relevant to interactions between equals (Rubin, Coplan, Chen, Bowker, & McDonald, 2011).

Social Monitoring as Social Support These four generic forms of social support are intuitively recognized and appreciated in everyday social relationships and constitute the foundation of most research portrayals of social support (e.g., House & Kahn, 1985; Thoits, 2011). But when research on social support is considered in relation to developmental psychopathology, at least one other function becomes important: social monitoring. Social monitoring emphasizes the protective influence of social network members, such as when friends notice signs of depression in an adult and seek assistance for this reason or extended family members detect and respond to disturbing bruises on a child. In contrast with the provisioning forms of social support earlier described, this is a protective function (House, Umberson, & Landis, 1988). It is an important, and sometimes overlooked, influence of social support and recognizes that support is important for preventing harm as well as for promoting good, and that both aspects are important to supporting individual well-being. Indeed, social monitoring may be most important, especially for some of the preventive and remedial influences that social support is believed to have in the lives of at-risk children and families, such as a home visitor's monitoring of an infant's development and of parental functioning. Furthermore, individuals and families in social isolation are commonly perceived as being at greater risk of psychosocial problems (such as child maltreatment) compared to those who are embedded in extensive social networks because the social monitoring of isolated families does not occur (Thompson, 1995). Although extensive networks do not necessarily safeguard against child maltreatment, as Korbin's (1991) research described earlier indicates, it improves the chances that child harm will be detected and that interventions on the child's behalf will be mobilized. Distinguishing the provisioning functions of social support—emotional encouragement,

counseling and guidance, access to information and resources, and skill acquisition—from the protective function of social monitoring is important for understanding why these do not necessarily operate well together. Stated simply, people are much more comfortable offering encouragement and help than with monitoring and drawing concerned attention to another's behavior, especially in informal social networks, and recipients are much more likely to respond positively to the former than to the latter (Thompson, 1995). In the families that Korbin (1991) studied, mothers were surrounded by affirmative and supportive kin who failed to intervene despite indications that a child was at risk. A similar conclusion derives from a recent study by Freisthler, Holmes, and Wolf (2014), who indicated that parents who experienced a high sense of belonging with others, especially when companionship involved drinking outside of the home, were more likely have physical abused their children. They described this phenomenon as “the dark side of social support” because the contexts of companionship may have fostered abusive conduct and friends did not help to curb this conduct. In a culture in which privacy boundaries, even within families, can cause adults to resent perceived intrusions into personal issues, these privacy norms can make neighbors, friends, coworkers, and even extended family members hesitate before bringing up uncomfortable concerns about personal conduct or parental behavior. Raising these issues can cause these individuals to be perceived as meddlesome, intrusive—and unsupportive. For most adults, formal helping professionals may be more successful than informal helpers in social monitoring and correction because of the nature of their role responsibilities. But even in formal contexts, these behaviors may be resented and resisted. This is, however, another manner in which social support to adults and children differs. There is much less perceived dissonance between the provisioning and protective functions of social support to children, in part because children's need for guidance is self-evident and privacy concerns do not apply as strongly to children's conduct (Smetana, 2006). In addition, both children and parents recognize that parental monitoring and correction are compatible with parental support and affirmation as long as parents are perceived as acting fairly and consistently with their parenting responsibilities (Grusec & Goodnow, 1994; Lansford et al., 2005). As Lansford and colleagues (2005) have noted, for example, even physical punishment by parents is accepted by children if this practice is perceived as normative. Thus although neither children nor adults like the attention drawn to their inappropriate or dysfunctional conduct, for children the correction of their conduct is a more normative feature of their relations with adults within and outside the family. Distinguishing different functions of social support in this manner draws attention to the different supportive functions that can exist in relationships between a support agent and a recipient. In some relationships, individuals are support specialists who are valuable primarily because they have one or two specific forms of social support to offer, and this can also be true of formal helpers, such as physicians, counselors, and religious leaders (Bogat, Caldwell, Rogosch, & Kriegler, 1985). Other individuals are support generalists who provide many different social support functions, and these individuals can often be found in neighborhood and kinship social networks. As noted earlier, parents are support generalists to children, a fact that helps to account for their unique and significant influence on children's

development (Cauce, Reid, Landesman, & Gonzales, 1990). But peers can also be support generalists of a different kind. Parker and Asher (1993) found, for example, that the friendships of children who were highly accepted among their peers provided greater validating and caring, more help and guidance, greater conflict resolution, and more intimate exchange than the friendships of children with lower peer acceptance. The extent to which children's developing social networks provide multifaceted support to them depends, in part, on the number of support generalists these networks include.

What Social Support Is and Is Not This multidimensional portrayal of social support includes the various ways that individuals derive interpersonal and material resources from others. It recognizes that social support extends beyond the emotional affirmation and understanding that is prominent in everyday conceptions of support and includes other contributions that are ultimately valuable to the recipient (although they may not always be immediately recognized as such). Viewed in the context of developmental psychopathology, this conceptualization recognizes that support can include skills training, community referrals, economic assistance, and social monitoring for purposes of preventing or remediating clinical problems. Indeed, these functions of social support may be influential whether they are accompanied by emotional affirmation or not. It is arguable, however, that this portrayal of social support is too inclusive. As earlier noted, for example, it is possible to view most forms of therapy as entailing one or more forms of social support. Any relationship that is worth maintaining likewise functions in a socially supportive manner in some form. In this light, it is perhaps less important to regard relationships as “supportive” than to delinate more specifically how they offer support and the particular support functions they provide. If the term social support is here conceived broadly, it nevertheless permits portrayals of helping relationships as functioning supportively in distinct ways. Moreover, it is important to clarify what social support is not. Social support is not the same as having a large social network. Individuals who know many people may not derive support from them—indeed, large social networks may be “conflicted networks” that impose stress more than support, and this may be especially true for troubled individuals (Barrera, 1981). Moreover, some people enjoy close, supportive relationships with a very small handful of network associates. Although these conclusions may seem obvious, they challenge the view that it is important to try to integrate socially isolated children or families with larger numbers of people with the expectation that doing so will enhance their social support. Without efforts to ensure that new relationships function helpfully in the lives of socially isolated individuals, simply enhancing network size may not accomplish much. For similar reasons, social support also is not social embeddedness (the frequency of contact with network members), dispersion (the ease with which individuals can get in contact with network members), or stability (the consistency of network members over time). Each of these structural features of social networks can be important in certain circumstances, and often these features are used as measures of social support. But research studies indicate that social

network measures like these are only modestly associated with the extent to which individuals report expecting, or actually receiving, social support from others (e.g., Barrera, 1986). One reason is that relationships with social network members can be troubled, coercive, or critical in ways that undermine the benefits these relationships potentially can provide. This is especially likely to be true for children at risk of psychological difficulty. These considerations thus focus the conceptualization of social support on the qualities of social transactions that matter to the recipient and are influential in promoting health. Subjective perceptions of perceived or enacted support from network associates are more strongly associated with individual coping and resilience than are structural features of social networks, such as network size (Barrera, 1986; Cohen & Wills, 1985). The association between individual coping and a subjective perception of support may be based on the sense that people are benefiting from their transactions with social network members in emotional, informational, material, protective, and/or other ways that are important to them. When this occurs, those social network members function in a socially supportive manner and usually are perceived to be doing so by the recipients. It also means, however, that individuals who act protectively in the lives of troubled families or children must carefully consider how their actions are perceived by recipients or risk that their assistance will be regarded as unsupportive, even if they are striving to act in the recipient's best interests.

Social Support and Stress This portrayal of the functions of social support embeds helping relationships within the everyday positive social transactions of children and adults. Social support benefits its recipients: It is part of what makes life manageable and cements close relationships. Supportive relationships are nice to have. With respect to children, however, Thompson (2014b) goes further to argue that social support is developmentally constitutive. In other words, beyond being desirable, the ways that supportive relationships (especially with parents) provide emotional security, foster the growth of age-appropriate skills and understanding, offer essential information and guidance on matters that are important to children, and offer protection through social monitoring address essential needs of developing children, particularly when they are young. They are not optional for healthy development. The view that social support is developmentally constitutive grows out of children's dependence on those who care for them, their limited capacity (until much older) to create and maintain their own supportive social networks outside the family, the centrality of caregivers in children's environment of relationships, and the importance of supportive relationships to critical developmental achievements in self-confidence, social and emotional competence, cognitive functioning, and even physical health (Thompson, 2014b). One implication of this view is that children's experience of social support is closely tied to the quality of the relationships they experience with family members, peers, and other important network associates. This concept is explored further in the next section. Another implication is that when these relationships are troubled, particularly with parents, children are especially vulnerable because alternative sources of support are not readily available to

them. This idea is explored further in the examination of social support in relation to developmental psychopathology. Social support is important, and it is especially so when individuals are under stress. Extensive research highlights that social support mediates the impact of stress on individuals in at least two major ways (Barrera, 1986; Cohen & Wills, 1985; House et al., 1988; Taylor, 2011; Vaux, 1988). First, social support can be stress-preventive. Social support provides its recipients with psychological and material resources that contribute to healthy functioning and positive development and in so doing prevents many stresses from occurring. This influence of social support has been studied most extensively in the field of health psychology, where social support is believed to encourage healthy practices through the encouragement and example of network members; the monitoring of healthy practices by others; the investment of the individual in roles and relationships that support healthy functioning; and encouragement of the individual's self-esteem, sense of belonging, and companionship (Thoits, 2011; Uchino, 2004; Umberson, Crosnoe, & Reczek, 2010). Viewed more broadly, social support can also promote mental and physical well-being by fostering needed competencies and skills, promoting effective coping strategies, and enabling access to assistance and material aid. These resources support the growth and maintenance of constructive practices and equip individuals with the capacities to manage ordinary challenges and prevent their escalation. In these ways, social support contributes to the prevention of stress. This influence of social support is also true of children in their social networks. Parents who are warm and nurturant, support children's self-confidence and self-esteem, encourage constructive social competencies, and foster personal and academic accomplishments help to prevent sources of stress for their children that can arise from difficulties in school or with peers or from problems of adjustment. Moreover, parents often enable access to other sources of social support (e.g., health care providers, teachers, extended family, coaches, mentors, and peers) that can have similarly constructive benefits. If social support is a protective factor, then it potentially can mitigate some of the risks that many children face in difficult or disadvantaged circumstances, but only if children experience support in their families and other relationships. Second, when stress occurs, social support can be stress-buffering. The psychological and material resources available through supportive relationships can diminish the impact of stressful events by enhancing constructive coping. In health psychology, clinical studies have shown that social support is associated with a reduction of the effects of disease pathology and psychological distress, contributing to less severe symptomatology, quicker recovery, enhanced coping, and diminished long-term sequelae (Cohen, 2004; Cohen & Janicki-Deverts, 2009; Umberson & Montez, 2010). This buffering can occur by altering recipients' appraisals of stressful events, enhancing their knowledge or utilization of effective coping strategies, providing useful information or instrumental aid, and supporting self-esteem, optimism, and perceptions of self-efficacy. These benefits are even more significant for children in light of their more limited capacities for coping with stress and their reliance on the assistance of

parents and other caregivers. More than 30 years of research have provided a fairly reliable account of the association of social support with stress, although there has been much less research confirming the hypothesized behavioral and psychosocial mediators of this association (Taylor, 2011; Umberson & Montez, 2010). This lacuna constitutes an important research agenda, especially with children in at-risk circumstances, because the hypothesized mediators of the preventive and buffering effects of social support may not regularly characterize the relationships on which children rely. Relationships may for them be sources of support but also of adversity, a topic that is discussed further in a later section.

Social Buffering of Biological Stress Reactions An important contribution to understanding the association of stress and social support comes from research on the neurobiology of stress reactivity and its social buffering. Taken together, this research contributes further evidence to the importance of social relationships in coping with stress, especially for children, for whom nurturant, supportive care shapes the functioning of biological systems relevant to stress responding. When acute stressors occur, a network of biological systems are activated in response to cortical appraisals of threat or danger (Lupien, McEwen, Gunnar, & Heim, 2009). One of the most important (and best studied) of these is the hypothalamic-pituitary-adrenocortical (HPA) axis, which, in concert with the sympathetic nervous system, releases the catecholamines epinephrine and norepinephrine and corticosteroids, such as cortisol, into the bloodstream. This activity contributes to the mobilization of energy, enhanced cardiovascular tone, and other critical components of the stress response that have psychological as well as physiological consequences. These responses are designed to provide short-term resources for coping with threat. When stress is chronic or recurrent, however, repeated HPA activation is costly. Chronic activation of the HPA system can change its neurocircuitry over time, contributing either to hyperreactivity to threat or hyporesponsiveness of HPA function, in each case reflecting dysregulated biological functioning (see reviews by Thompson, 2014a, 2015b). In addition, chronic release of cortisol affects the functioning of limbic systems (such as the amygdala, hypothalamus, and hippocampus) and cortical systems (including prefrontal functioning) that can alter emotional reactivity, cognitive functioning, and self-regulation (Ulrich-Lai & Herman, 2009). Chronic stress also alters autonomic functioning, suppresses immune response to infectious challenges, and generally embeds proinflammatory tendencies into biological functioning (El-Sheikh & Erath, 2011; G. E. Miller, Chen, & Parker, 2011). The biological consequences of chronic stress are consistent with the concept of allostatic load: the progressive wear and tear on biological systems attributable to the long-term effects of chronic stress, which elevates risk for later health and mental health problems (Danese & McEwen, 2012). The behavioral consequences can include poorer emotion self-regulation, cognitive and attentional problems, and difficulties in social functioning (Blair & Raver, 2012; Evans & Kim, 2013). These considerations are particularly serious when children are stressed because of the

plasticity of developing neurobiological systems and their sensitivity to experiences of stress. Fetal exposure to maternal stress hormones is associated with altered stress reactivity after birth. In one prospective longitudinal study, for example, maternal depression during pregnancy was associated with heightened cortisol levels when infants were observed at 3 months in a moderately challenging procedure (Oberlander et al., 2008). In another longitudinal study of children in poverty, environmental characteristics such as poor housing quality, family economic strain, and poor parenting were associated with disrupted HPA activity from age 7 months to age 4 (Blair et al., 2011). Yet in the same sample, observed maternal sensitivity from birth to age 2 buffered some of the effects of family stress on children's cortisol reactivity (Hibel, Granger, Blair, Cox, and the Family Life Project Key Investigators, 2011). The latter finding is consistent with the view that social support can buffer biological stress reactivity in children and adults (Hostinar, Sullivan, & Gunnar, 2014). Studies with adults show, for example, that social contact with a supportive person can reduce cardiovascular and HPA reactivity and reduce threat-related neural activation to a stressful task (Taylor, 2011; see, e.g., Coan, Schaefer, & Davidson, 2006). In research with children, a growing literature finds that a secure attachment has comparable physiological stress-buffering consequences, especially for children who are prone to fear or inhibition (Gunnar & Donzella, 2002). Nachmias, Gunnar, Mangelsdorf, Parritz, and Buss (1996) found, for example, that behaviorally inhibited toddlers exhibited significant cortisol escalation when faced with a variety of novel and challenging events in the company of their mothers, but only when they were insecurely attached to her. For inhibited toddlers in secure mother-child relationships, cortisol response to the same events was more moderate, comparable to that of uninhibited children. These studies underscore the importance of relationship quality to the social buffering of stress. Warm and supportive relationships are more likely to moderate stress reactivity than are those that are less affirmative or aversive, and this is true for adults as well as children (Taylor, 2011). However, owing to the plasticity of developing neurobiological stress systems, this relational buffering may have both immediate effects and long-term consequences for children's stress reactivity. This conclusion is suggested by the programmatic studies of Meaney and his colleagues on biological stress responding in rat pups (see Meaney, 2010, for a review). They have shown that differences in maternal licking and grooming (LG), an important component of maternal nurturance, are associated with differences in stress reactivity in pups, with the offspring of high-LG mothers showing lower behavioral and neuroendocrine responses to stress compared to the offspring of low-LG mothers. In the short term, maternal LG is a source of tactile stimulation that modulates endocrine and cardiovascular functioning in the pup and regulates immediate stress reactivity. Over the long term, however, the pups of high-LG mothers also show more modest behavioral and endocrine responses to stress as well as stronger performance in learning and memory tasks and greater synaptic density in the hippocampus. By contrast, the offspring of low-LG mothers show greater efficiency of fear conditioning in high-stress contexts. Cross-fostering studies indicate that these influences arise from early experience, and Meaney's research group has documented the biochemical processes by which maternal LG stimulates epigenetic changes in

glucocorticoid receptor gene expression that relates to these behavioral outcomes associated with stress reactivity and early cognitive functioning (Meaney, 2010; see also Suomi, 2011, for comparable findings on the social buffering of stress in rhesus monkeys). Social support also moderates the behavioral manifestation of genetic characteristics that, in the context of stress or danger, are likely to be risk factors for developing psychopathology. Bakermans-Kranenburg and van IJzendoorn (2006) studied this gene–environment (G×E) interaction in a sample of very young children and their mothers. They observed maternal sensitivity when infants were 10 months old, measured children's externalizing problems at 39 months, and genotyped children to identify those with the DRD4 7-repeat polymorphism, which has been associated with novelty seeking and conduct problems. Children with the DRD4 7repeat whose mothers were insensitive were high in later externalizing, but children with same genetic profile with sensitive mothers showed much less externalizing behavior, comparable to children without the DRD4 7-repeat. In this case, the support afforded by a secure attachment buffered the behavioral risk associated with this gene polymorphism and, in fact, contributed to more optimal later behavior, consistent with the formulations of differential susceptibility theory (Belsky & Pluess, 2009). An intervention study subsequently enlisted this G×E interaction in an effort to improve developmental outcomes (Bakermans-Kranenburg, van IJzendoorn, Pijlman, Mesman, & Juffer, 2008). A video-feedback parenting program was provided to a sample of mothers who had rated their toddlers high on externalizing behavior. A posttest after 1 year revealed that the program succeeded in improving mothers' positive disciplinary practices. In a follow-up assessment 1 year later, the children of the intervention group showed less externalizing behavior, but this finding was significant only for children with the DRD4 7-repeat polymorphism. There were no intervention program–related changes for children without the 7-repeat allele. The benefits of improved social support, through more sensitive parenting, were apparent for children who were at greatest risk of adverse developmental outcomes and who showed the greatest improvement as the result of the parenting program. Illustrations of G×E interaction involving social support interventions for children under stress are evident in other research. Kaufman and her colleagues (2004) showed that for children with a history of maltreatment and a genetic vulnerability to depression, those without social support showed elevated depressive symptomatology, more than twice the rate of nonmaltreated comparison children with the same genotype. By contrast, those reporting access to social support exhibited only minimal increases in depressive symptomatology compared to nonmaltreated children. In other research, Luby and her colleagues (2012) studied school-age children with or without a history of preschool depression, and focused on hippocampal volume, which is smaller in adults with severe depression. Maternal support observed when children were preschoolers was positively associated with hippocampal volume at school age for nondepressed children. What neurobiological influences can account for the immediate and long-term benefits of social support as a stress buffer? The biological mediators identified in this research complement, to some extent, the behavioral explanations discussed earlier, which focus on

changing appraisals, strengthening coping, and supporting self-esteem. Hostinar et al. (2014) argue that social support (especially from caregivers) strengthens the regulatory influences on HPA activity arising from (a) the oxytocin system and (b) prefrontal activation, based on research concerning the neural regulators of HPA activity and oxytocin as a socially elicited stress buffer (Ulrich-Lai & Herman, 2009). This formulation argues that stress buffering consists not only of the downregulation of HPA activity and its associated sequelae but also the upregulation of oxytocinergic hormonal influences and cortical controls that become primed by supportive social contact. Social support also contributes to downregulation of many of the proinflammatory tendencies associated with chronic stress, although it is unclear whether support builds protective factors into the immune system (Kiecolt-Glaser, Gouin, & Hantsoo, 2010). These biological formulations certainly merit further exploration. Unfortunately, it must be remembered that just as social support can biologically remodel stress reactivity over time, chronic stress can have comparable consequences. As earlier noted, these consequences include effects on the neurocircuitry of stress reactivity, downstream consequences for the neurobiology of learning and memory, and immune suppression. In addition, chronically stressful experiences are associated with epigenetic changes in gene expression that dysregulate stress reactivity, immune function, and neurological development (Naumova et al., 2012; see Thompson, 2015b, for a review). In these studies, parental care is the major determinant or mediator of early life stress. Children growing up in conditions of family adversity, which can range from harsh or insensitive care to maltreatment, are likely to suffer the biological dysregulation arising from these experiences in ways that are behaviorally manifested in heightened vigilance and threat detection, emotion dysregulation, chronic health problems, and risk for affective and conduct disorders (Repetti, Robles, & Reynolds, 2011). These biological programming effects can make behavioral remediation much more difficult, and they underscore the developmentally constitutive influences of supportive care. The research on the social buffering of stress, especially newer work on its biological underpinnings, highlights the significance of the quality of parental care to children's biological and behavioral development by shaping capacities for stress regulation and coping (Thompson, 2015b). It indicates that the quality of care can have significant consequences—positive or negative—that can have a snowball effect in children's developing abilities to cope with stress and challenge, function in a self-regulated manner, and mobilize cognitive skills for learning and achievement.

Stress and Access to Social Support In the paradigmatic portrayal of social support that guides thinking about support as a stress buffer, assistance begins when one or more individuals discern that a person or a family is in crisis. They offer to help and also mobilize others in the recipient's social network to contribute. In a short time, that person or family is surrounded by solicitous acquaintances offering emotional support, material resources, guidance and access to information sources, assistance with life tasks, or anything else they think might be of value and benefit to the recipient. Although it is often impossible to ameliorate the circumstances that led to crisis, this kind of assistance can give recipients a sense that they are not alone, provide tangible help

where it is needed, and temporarily lift the burden of other responsibilities, and it is welcomed for these reasons. In some circumstances, furthermore, formal helping agents—such as religious advisors, social workers, counselors, physicians, or psychological therapists—also may be enlisted to provide professional and targeted kinds of support. The association among social support, stress, and psychological well-being is complicated, however, because stress often is associated with the deterioration, rather than enhancement, of social support for recipients (Shinn, Lehmann, & Wong, 1984; Vaux, 1988). Stress can undermine social support in at least two ways. First, stress can cause potential recipients to withdraw rather than engage with other people who might provide assistance. Withdrawal can occur because of the circumstances contributing to difficulty, such as how the loss of a job, divorce, residential relocation, and other events can separate individuals from familiar social networks. It can also occur because of the recipient's humiliation or feelings of vulnerability, which can accompany stressors such as mental health problems, unexpected financial difficulty, or other serious family problems. This perceived vulnerability can motivate individuals not only to avoid contact with social network members but also to avoid seeking the assistance of formal help providers who might offer tangible assistance. Withdrawal can also result from the recipient's deliberate efforts to conceal the reasons for current difficulties deriving from, for example, psychological distress, domestic violence, or child maltreatment. This withdrawal can be one of the reasons for the social isolation of families in which children are being abused or neglected (Polansky, Chalmers, Buttenwieser, & Williams, 1981; Powell, 1979; Young, 1964). Finally, individuals in crisis may simply experience too much emotional turmoil to seek or accept much assistance from others. A second reason that stress can contribute to diminishing social support is because of the effects of stressful circumstances on potential providers of support. Individuals in natural social networks may withdraw from potential recipients because the behavior of the recipient is disturbing or repelling. Sometimes this is the response to neglectful or abusive parents, for example, or individuals with serious mental health problems. Withdrawal may also occur over time because the recipient's problems are emotionally challenging or overwhelming, such as when helpers try to support friends who are coping with the loss of a child or suffer from terminal illness. Sometimes potential support providers withdraw because they are undermined by circumstances that are similar to those of potential recipients, such as when neighbors live in poverty or face the challenges of single parenting (e.g., Belle, 1982; C. S. Fisher, 1982; Wortman & Lehman, 1985). More generally, potential support providers may withdraw from troubled individuals because of the emotional drain of providing one-way assistance to very needy individuals (J. D. Fisher et al., 1982; Shumaker & Brownell, 1984). The self-protective withdrawal of potential helpers derives from the difficulty of providing meaningful assistance to individuals who need so much. When formal helping agents are concerned, professional training and experience buffer these provider reactions to assistance for stressed individuals. Nevertheless, effective supervisory and supportive professional associations are necessary to prevent burnout of these providers also. In everyday circumstances, these causes of diminishing access to social support can be overlapping and interactive. Young children who are physically maltreated may be prevented

from gaining access to needed support because of the impact of their abuse on social skills, affect regulation, and expectations for how others will respond to them, which undermine how they function in the adult and peer relationships that might provide support (Cicchetti & Valentino, 2006). Abusive parents may, in turn, actively isolate the child from others in the community to prevent the detection of abuse, and other adults may distance themselves from a family that seems troubled or disturbed. Childhood and adult depression also shows these cumulative isolating processes (Cicchetti & Toth, 2006). In adults, the anhedonia of depressive symptomatology is typically accompanied by social withdrawal together with irritability and hostility, passivity, self-denigration, dependency, and a sense of helplessness and hopelessness. The emotional demands of this symptom profile on close relationships can cause others to withdraw and avoid contact, which, in turn, confirms the depressed person's perceptions of being rejected by others and of the unreliability of those relationships (Coyne, Burchill, & Stiles, 1991; Coyne & Downey, 1991). Adult depression is thus associated with smaller social networks, fewer close relationships, and diminished perceptions of others' support by the depressed adult, and the same associations have been found for children and youth. Dubois, Felner, Brand, Adan, and Evans (1992) found, for example, an expected negative association between social support and psychological distress 2 years later in young adolescents, with initial levels of distress and other sources of stress controlled. They also found, however, evidence that distressed youth acted in ways that inadvertently reduced their access to social support and increased the stressfulness of daily experiences. In short, the association between social support and stress is complex. Although support can contribute to preventing stressful events and there are psychological and biological reasons to regard support as a stress buffer, there is also a reciprocal association between social support and stress. Because of the consequences of stressful circumstances for potential recipients and on potential helpers, increased stress can diminish access to social support and leave individuals more isolated than they were before. These considerations caution against simple, paradigmatic portrayals of how support affects stress and contribute to the recognition that the effective mobilization of social support for individuals in need, from both informal and formal sources, must consider the nature of the stress, its impact on both recipients and support providers, and the context of their interaction.

Social Support in Relationships and Social Networks An important context of social support consists of the relationship between recipient and support provider and the social networks that encompass them both. To be sure, the growth of social media, online social networks, the Twitterverse, blog posts, even virtual gaming (some with avatars in simulated societies) and a growing range of technological advances enable children and adults to contact and develop associations with potentially hundreds of other people without direct face-to-face contact with them. Moreover, these associations can provide some forms of social support, especially sympathetic responses to an individual's posts, and we shall return to these considerations later in this section. But there are many reasons to argue that the most beneficial and comprehensive forms of social support,

especially for stressed individuals, arise from relationships that involve mutual understanding, direct give-and-take, and emotional connectedness emerging from a history of reciprocated assistance. Most often these interactions occur in the context of social networks with distinct characteristics and affiliative connections. Individuals who are capable of providing diverse and complex forms of social support, mobilized by intimate understanding of the recipient and the recipient's problems, are most likely to be in real, not virtual, social relationships and social networks. Yet it is also true that the characteristics of those relationships and the networks in which they are embedded may pose obstacles as well as benefits to the provision of social support. This section explores these considerations with respect to both informal and formal sources of support. After considering the characteristics of relationships that contribute to perceptions of social support, the discussion turns to characteristics of the social networks that influence how relationships function and the similarities and differences between the social networks of parents and children. We then return to the question of online social support.

The “Active Ingredients” of Supportive Relationships A relationship can be defined as an integrated network of enduring emotional ties, mental representations, and behaviors that connect one person to another over time and across space (Laible & Thompson, 2007). Why are relationships important to social support and promoting individual well-being? When relationships function well for adults and children, they uniquely offer psychological resources that are essential to perceptions of support that contribute to reducing stress and enhancing well-being. With respect to children, Thompson (2014b) has identified four characteristics of relationships in children's central social networks—family and peers—that are important to perceived support from people within them. Mutual responsiveness/reciprocity is based on the reciprocated affinity by which relationships are formed and consolidated and which contributes to a partner's perceived support. In infancy, the parent's sensitive responsiveness contributes to the development of a secure attachment as well as cognitive and linguistic achievements that are based on the reciprocated give-and-take of social interaction (Laible & Thompson, 2007). The parent's responsiveness also provides the basis for developing mutual responsiveness at somewhat older ages that supports early socialization (Maccoby, 1984; Waters, Kondo-Ikemura, Posada, & Richters, 1991). In peer relationships, reciprocity is what distinguishes friends from nonfriends: In friendship, there is mutuality of affection, intimacy, interests, initiative, and other aspects of relational interaction (Newcomb & Bagwell, 1995). When asked “What is a friend?” children of a variety of ages talk about the reciprocity or mutual giveand-take of interaction with friends (Hartup & Stevens, 1999). Warmth/affection define the affective quality of relationships. In parent-child relationships, warmth conveys love and respect for the child and, as it elicits a complementary response in children, mediates the efficacy of other parenting practices that promote cooperation and competence (Kochanska, 2002). In peer relationships, mutual affection is an important characteristic of friendship even though friends also exhibit greater conflict than nonfriends do (Simkins & Parke, 2002). Their greater mutual liking is part of what motivates friends to resolve conflict in order to remain in the relationship

rather than walking away. Approval/acceptance constitutes the positive regard that supports the recipient's selfconfidence and perceptions of a partner's liking and support. From parents, approval contributes to a sense of competence in children and adolescents and contributes to achievement motivation in this way (see, e.g., Pomerantz & Dong, 2006). Peer acceptance (versus rejection) is a major contributor to developing self-esteem, social skills, and social understanding as it strengthens children's constructive peer interactions (Bukowski, Brendgen, & Vitaro, 2007). Security, in relationships with family members and peers, is a contributor to and a derivative of perceived support. A large research literature in parent-child attachment documents the significance of early relational security for social and emotional well-being (Thompson, 2008), while security in friendship networks contributes to many positive outcomes, including school adjustment and academic achievement (e.g., Berndt, Hawkins, & Jiao, 1999). These four characteristics of children's relationships with parents and peers contribute to children's perceptions of support from these social partners, and these characteristics are also important to adult relationships. Even though social support is multifaceted, as earlier argued, these relationship characteristics are significant because perceived support is the dimension of social support that is most strongly related to psychological well-being in adults and children (Barrera, 1986; Cohen & Wills, 1985; Jackson & Warren, 2000; Sarason, Shearin, Pierce, & Sarason, 1987). Perceived support is more predictive of coping and well-being than even enacted support. This finding is important because, surprisingly, there is not a strong association between perceived and enacted social support (Barrera, 1986; Haber, Cohen, Lucas, & Baltes, 2007). To be sure, the relation among enacted support, perceived support, and coping is empirically complicated, as the preceding discussion of the association between social support and stress illustrates (interpreting the correlations between these constructs is complicated by (a) the recognition that individuals under stress may be less capable of viewing others as sources of available support because of their emotional turmoil, (b) individuals in difficulty may be less capable of mobilizing supportive social networks when they are needed, (c) individuals can enact support without being perceived by recipients as doing so, and (d) individuals who experience psychosocial well-being are more likely to be high in perceived support even though there is no causal connection between them (see, e.g., Kaul & Lackey, 2003; Seagull, 1987)). Nevertheless, qualitative and quantitative reviews consistently find that perceived support has a stronger predictive association with psychological well-being than does enacted support. The significance of perceived support underscores the importance of confidence in the availability and helpfulness of social partners, whether or not that support is actually realized, to an awareness that assistance is available if needed. This is, in a sense, what is meant by a secure (attachment) relationship at any stage of life. More broadly, Lakey and his colleagues (e.g., Lakey & Lutz, 1996; Rhodes & Lakey, 1999) have argued that perceptions of social support derive from (a) personality characteristics of the recipient (e.g., a secure or insecure

attachment history; interpretive biases, extraversion), (b) characteristics of the helper (e.g., personality factors like empathy; enacted support), and, most important, (c) the interaction between helper and recipient (e.g., their similarity in background and outlook). This heuristically powerful analysis invites broader inquiry. The significance of perceived support also underscores the importance of mental representations of relationships to developing and maintaining psychological well-being (Thompson, 2014b). Mental representations of network associates as being supportive, accessible, reliable, and helpful are based on direct experience with them, and these representations may be more important than specific acts of support to promoting coping and well-being. With respect to social support and developmental psychopathology, this conclusion has important implications. It suggests that attention to these relational representations, in children as well as adults, may be an important element of preventive and therapeutic intervention. It may not be enough, for example, to change the behavior of a parent who has been an inadequate caregiver without also altering the child's expectations for the behavior of that adult in the future (see, e.g., Toth, Maughan, Manly, Spagnola, & Cicchetti, 2002; Toth, Rogosch, Sturge-Apple, & Cicchetti, 2009). Similarly, altering parental expectations for child conduct may be essential, along with changing damaging behavior patterns, to effecting longstanding improvement in a dysfunctional family system (see, e.g., Snyder, Cramer, Afrank, & Patterson, 2005).

Relationships and the Social Networks of Parents and Children Relationships do not always function well for children and adults. Indeed, the psychological complexity of relationships and their multifaceted influences often cause them to be simultaneous sources of support and stress (Badr, Acitelli, Duck, & Carl, 2001). Equally importantly, the nature of the relationship helps to determine what kinds of support are possible and the limitations that may exist in giving and receiving social support. This is true of relationships with informal social network members and with formal helpers. The informal social networks that family members depend on are both local and geographically disparate (Cochran, Larner, Riley, Gunnarsson, & Henderson, 1990; C. S. Fischer, 1982; Litwak & Szelenyi, 1969). They can include extended family members, friends who live nearby or far away, current and former neighbors, associates in the religious community, and networks based on children's peer relationships and school activities. For parents, their social networks are also likely to include colleagues at their workplace or school and perhaps individuals in local businesses and community leaders. For children, their social networks include peers and teachers at school and the variety of other adults involved in after-school programs, recreational and sports programs, and other activities; when adolescents are concerned, they may also include workplace associations. For both parents and children, the breadth of their natural or informal social networks is influenced by access to transportation and technology—each of which can significantly extend their networks of relational ties—and the availability of time to cultivate relationships with others. Informal social networks have changed historically in their breadth and diversity, and also vary according to socioeconomic status and other circumstances (Cochran et al., 1990).

There is high turnover in natural social networks, but turnover is selective. In a 3-year followup study of the personal and familial social networks of a sample of 240 parents living in Syracuse, New York, nearly one-fourth (22%) of the network members on average were dropped from adults' social networks over 3 years between the first and follow-up interviews (Larner, 1990a). When researchers examined the nature of the social relationships that were dropped or retained, they found the highest turnover among neighbors (32% turnover); friends, acquaintances, and former workmates (31% turnover); and other nonkin ties (34% turnover). The lowest turnover rate was among kin: only 9% turnover. This finding is not surprising, given the extent to which high rates of mobility, job change, educational completion, and other factors can alter neighborhood and workplace social networks in contrast with the stability and commitment to kinship ties that motivate efforts to remain in contact, even across a distance. These findings suggest, however, that the relative reliability of network contacts over time can influence those from whom one seeks and receives social support. Consistent with this view, researchers have found that individuals rely primarily on support from extended kin within their natural social networks (Cochran, Gunnarsson, Grave, & Lewis, 1990; Gunnarsson & Cochran, 1990; Litwak & Szelenyi, 1969). By contrast, neighbors do not figure prominently in perceptions of social support and were readily dropped or added to personal social networks with little influence on individual or familial social support (Larner, 1990b). The turnover in neighborhood and other local associates makes depending on them as reliable sources of assistance more difficult, while the enduring, obligatory relationships that develop with extended family members provide a more consistent basis for expectations of support. Indeed, when a person is reliant primarily on neighborhood contacts for support, it may be an indicator of social insularity in other spheres of life. In a study of socioeconomically stressed single mothers, for example, Belle (1982) found that the reason these women interacted more frequently with people in the neighborhood was that they felt they had nobody else to whom they could go. However, these contacts did not help them much. The mothers who were in frequent contact with neighborhood network associates did not experience greater emotional well-being, receive more tangible assistance, or report more support than did other women who saw their neighbors infrequently, perhaps because these neighborhood associates provided social contact but not social support (see also Belle & Benenson, 2013). The turnover in informal social networks and the distinction between kin and nonkin ties is important to the preventive and therapeutic mobilization of social support. It suggests that a focus on strengthening neighborhood-based forms of social support—“neighbors helping neighbors”—must balance the potential benefits of such a strategy against families' limited reliance on neighbors owing to residential mobility. The recommendations to Congress of the U.S. Advisory Board on Child Abuse and Neglect (1991, 1993) to encourage the development of “caring communities” by strengthening supportive connections among neighbors builds on their proximity and the commonality of neighbors' interests with those of recipients, as well as the potential utility of such a strategy in communities across the socioeconomic spectrum. Neighbors can also provide referrals to local help givers, material aid, emergency assistance, and respite child care—in short, they can be multifaceted support generalists. But if residential

mobility undermines the ability of family members to even become well acquainted with their neighbors, it is also likely to undermine perceptions that social support might be obtained from them. Furthermore, in communities at risk, neighbors may experience the same economic and personal problems as do the families who need their assistance, and neighbors may have little to offer because of their own needs. Neighbors may be more concerned about finding a quick exit from a dangerous or impoverished community than with providing support to a troubled family who lives nearby. Kinship ties, however, also have advantages and disadvantages in the provision of social support. An extended history of shared experiences contributes to common values and mutual understanding. Family relationships are densely interconnected, with family members often knowing many other family members well, which contributes to the mutuality and intercoordination of support. Cultural expectations and practices concerning kinship also better enable family members to maintain connections through formal practices (such as celebrating holidays together and recognizing birthdays and anniversaries) and informal occasions, even though extended kin may be geographically distant. But family relationships can also be characterized by animosity, criticism, privacy intrusions, or distrust that also endures. Moreover, the shared history and values that characterize family relationships can be an impediment to healthy behavior if those values condone or rationalize abusive conduct toward children or partners, substance abuse, poor health or relational practices, or other dysfunctional individual or family behaviors. Family members may be more motivated to deny or conceal these practices if exposing them risks bringing these and other family practices to the attention of others, including clinical or criminal justice professionals. Thus, families can become enabling environments for developmental psychopathology rather than settings of health and healing. The different constellations of strengths and weaknesses of relationships with neighbors and extended family are important because each kind of network connection is important to informal social support. Just as parents are support generalists to their children—providing many different kinds of social support—so also are family members generalists in support to each other (see, e.g., Bogat et al., 1985). To a lesser extent, neighbors can also be support generalists in the various kinds of assistance they can potentially provide to those who live nearby. This is one of the reasons why kinship networks and neighbors have often been identified as primary sources of social support that can be mobilized for preventive or therapeutic intervention. But as this analysis suggests, there should be careful consideration of the advantages and disadvantages of mobilizing social support from each source. The same balancing of supportive benefits and risks characterizes other relationships within informal social networks. Family friends and workplace associates can offer counseling and advice, but their help may be limited to the domain of common association (e.g., workplace concerns, shared recreational interests, connections through children's peers at school), and they may otherwise have less in common with the recipient. The same may be true of associates in religious institutions or community groups, and in these cases the organization's values or mission may additionally constrain disclosure or the kinds of assistance that can be offered. In general, people in these relationships can be characterized as support specialists

who can be helpful with specific kinds of supportive assistance, but not more broadly. The same situation specificity of relationships is also true of formal support agents (Unger & Powell, 1980). A social worker may offer valuable assistance by connecting individuals or families to community resources and providing informal counseling. But an overwhelming caseload may diminish the caseworker's reliability and limit the extent of the assistance that can be offered. Similar limitations may also be true of physicians, counselors, or religious leaders, who may also view problems through the prism of their specialized background, which might compel them to offer forms of assistance that may or may not really be needed. Their support may be limited by other considerations as well. It is well known, for example, that professionals who are legally mandated reporters of suspected child maltreatment are often aware of many more cases of suspected abuse than they formally report. This fact owes, in part, to the complications introduced into their professional relationships with families once a child abuse investigation has been initiated, and the preference of many helping professionals to address family problems on their own (Zellman, 1990; Zellman & Anter, 1990). In this case, and in others, their professional role may complicate their assistance. In general, professional helpers have expertise, the benefits of a clear role definition in relation to recipients, and professional training and accountability that can enhance the support they offer. But their influence can be limited by the specific forms of assistance they are capable of offering, the fact that their assistance is not well integrated into the everyday experience of the recipient, and sometimes because their values and perspectives are different from those of the recipients of their assistance. Although the nature and functioning of informal and formal social networks share many commonalities across subcultural and cultural groups, there are also important variations that are relevant to social support (Cochran, Larner et al., 2002; Vaux, 1985). In general, for example, the individual and familial social networks of lower-income families tend to be smaller and more kinship-based than for middle- and upper-income families (Cochran, Gunnarsson et al., 1990; Fischer, 1982; Vaux, 1988). In one large study, these differences by size were apparent for each category of social network membership: Higher-income families had a larger number of social ties with kin, neighbors, and “others” than did lower-income families (Cochran, Gunnarsson et al., 1990). Single mothers also have smaller social networks with higher rates of turnover compared to married mothers, and these networks afford less social support (Larner, 1990a; Weinraub & Wolf, 1983). Taken together, these findings point to the importance of differences in the resources, energy, and opportunities to construct and maintain the social networks that afford social support. Unfortunately, for parents and families who are most likely to need support, their networks often may be quantitatively and qualitatively less capable of providing it.

Development and Functioning of Children's Social Networks Children's social networks are, not surprisingly, similar to those of their parents because they are built on them. But they are also very different in several important ways (Belle, 1989; Cochran & Brassard, 1979; Salzinger, Antrobus, & Hammer, 1988).

Children's social networks evolve developmentally as a result of children's growing social competencies and their expanding social ecology. Parents are uniquely central to children's initial social networks and mediate children's relationships with other social partners. This is, of course, true from birth. The security of an infant's attachment to each parent derives from representations of support from that adult based on the parent's history of sensitive responsiveness. One reason why the development of secure or insecure attachment early in life is developmentally formative is that it influences the growth of other central socioemotional capacities, including the development of social skills, self-regulation, self-concept, and representations of other people and how to get along with them (Thompson, 2008). One implication of this conclusion is that in this earliest experience of reliable parental support (or its absence), children's capacities to elicit social support from other partners are shaped. In a sample of lower-income African American 6.5-year-olds, for example, Anan and Barnett (1999) found that secure mother-child attachment (assessed 2 years earlier) was associated with children's perceptions of greater social support from various sources (including family, peers, and teachers), and social support mediated the association between secure attachment and lower scores on measures of externalizing and internalizing problems. Bost, Vaughn, Washington, Cielinski, and Bradbard (1998) likewise found that secure preschoolers had more extensive and supportive social networks and were also higher on sociometric assessments of peer competence. These and other findings can be understood as reflecting the more constructive social skills and more positive social expectations that are associated with an early secure parent-child relationship. But they also reflect the snowball effect that can derive from the supportiveness of early parental care, which fosters a range of social and emotional competencies, which then contributes to children's developing capacities to elicit support from others. In this respect, children's experiences in early relationships affect other relationships. Throughout the world, it is rare that the social networks of very young children are exclusively parent-focused. In many cultures, for example, caregiving is shared with other adults, including but not limited to extended family members, to whom infants and toddlers also form attachments (Hrdy, 2009). In the United States and other Western industrialized countries, child care practices also include family-based and center-based out-of-home care settings and early childhood education programs. These and other normative child care practices provide early avenues for broader sources of social support to children as well as their parents. Extrafamilial caregivers can offer parents guidance and knowledge, respite care assistance, and referrals to others who can offer help and can monitor the child's well-being. These resources depend on the nature and the quality of the care, of course, and these caregivers can also offer intervention avenues for preventive or therapeutic assistance (Thompson, Laible, & Robbennolt, 1997). Many child care and preschool programs, for example, have created avenues of mental health consultation that provide teachers access to an early childhood mental health practitioner for advice concerning the behavior of children in the classroom or general guidance concerning the needs of children encountering stressful circumstances (Johnston & Brinamen, 2006). In the context of consultation, parents are often engaged to better understand children's needs and to collaborate in the design and implementation of an intervention strategy. Early childhood mental health consultation illustrates one way of enlisting extrafamilial care settings as avenues of support to children and their parents. It also

exemplifies the coordination of formal and informal sources of social support to families, as well as the value of two-generation interventions on behalf of young children. School entry in middle childhood further expands children's extrafamilial social networks by providing access to a more extensive range of peers as well as with adults who, as teachers or school counselors, can potentially provide advice, counseling, emotional support, and other forms of needed assistance. Teachers can potentially become important sources of social support because they, like early childhood care providers, may be the first to identify the seriousness of a child's problems and create a bridge between the family and formal helpers (Meehan, Hughes, & Cavell, 2003). In adolescence, workplace and community associations further broaden access to potential sources of social support as youth become more independent of family networks. Throughout childhood and adolescence, peers become more important sources of social support as peer relationships increasingly draw on and advance children's developing psychological and emotional sensitivity to other people. Viewed in this light, positive peer relationships can be stress buffers for children, while peer rejection and relationships with deviant peers may contribute risk for behavioral and emotional problems. One large, prospective longitudinal study that followed over 500 randomly selected children from preschool through early adulthood found that, as early as preschool, exposure to aggressive peers predicted later aggressive behavior (Sinclair, Pettit, Harris, Dodge, & Bates, 1994). In another longitudinal study, Gazelle and Ladd (2003) found that the combination of anxious solitude (an index of individual vulnerability) and peer exclusion predicted levels of depressive symptoms in a sample of 388 children studied from kindergarten to fourth grade. But there is evidence for the stress-buffering effects of positive peer relationships as well. In the prospective longitudinal study by Dodge and colleagues just described, peer acceptance moderated the effects of low socioeconomic status, high family stress, single-parent status, and violent marital conflict on the development of externalizing behavior in early grade school (Criss, Pettit, Bates, Dodge, & Lapp, 2002). In the same study, peer acceptance and friendship moderated the effects of harsh discipline on the same outcomes. These positive effects of peer relationships remained even when the researchers controlled for child temperament and social information processing skills. Friendship is an especially important avenue of social support, although children and adolescents also derive significant benefits from more general peer acceptance. Friendship can contribute to enhancing self-esteem and positive self-evaluation, promote emotional security, offer a nonfamilial context for intimacy and affection, provide informational and instrumental assistance, and offer companionship (Parker, Rubin, Erath, Wojslawowicz, & Buskirk, 2006). In this respect, friends also constitute support generalists. These peer influences are potentially enduring. A longitudinal study found that peer group acceptance and friendships in the fifth and sixth grades significantly predicted adult life status, perceived competence, and psychopathology 12 years later (Bagwell, Newcomb, & Bukowski, 1998). In this study, early peer group acceptance and friendship made unique contributions to adult well-being: Having a reciprocal friendship in childhood was associated with adult self-worth, for example, even when preadolescent levels of self-competence were controlled. With increasing age,

adolescent and early adult peer associations become more extensive (encompassing workplace and recreational connections as well as school relationships), more intimate (incorporating self-disclosure and loyalty concerns), and more multidimensional (including romantic as well as companionate relationships). Although increasing autonomy from the family system is a developmentally appropriate transition of adolescence, parents' continuing importance as potential social support agents should not be underestimated. Even as peer relationships become more important forums for self-disclosure and mutual understanding, parents remain preferred advisors on core moral values, political and religious beliefs, and planning and achieving life goals (Coleman, 2011). With increasing age, however, there may be important changes in children's attitudes toward seeking parental support. While seeking emotional reassurance or instrumental aid is natural and encouraged in younger children, adolescents may resist help-seeking if doing so is regarded as a threat to self-efficacy or perceived competence (Robinson & Garber, 1995). The simultaneous reliance on parental help combined with the increasing need for perceived selfreliance contributes to the mixed signals of adolescents and the mixed perceptions of their parents concerning their support needs. There are developmental changes in how children perceive support from different people in their social networks (Furman & Buhrmester, 1985, 1992; Reid, Landesman, Treder, & Jaccard, 1989). In self-report studies, adolescents report expecting less support from their parents, for example, than do younger children, and there are similar developmental decreases in perceived support from other family members, such as siblings and grandparents. Expectations of support from peers increase from childhood to early adolescence but then stabilize or decline in later years. Teachers, by contrast, are rarely regarded as sources of social support. Children's social networks are therefore initially based on parents and their associations but progressively expand to encompass partners and contexts that are unique to them. But parents also mediate children's access to sources of social support, both formal and informal, in other ways (Cochran & Brassard, 1979; Parke & Bhavnagri, 1989). Parents' decisions concerning housing, neighborhoods, and schools directly affect the range of children's options for developing social connections outside the family. Ladd, Hart, Wadsworth, and Golter (1988) found that the breadth of preschoolers' social networks was inversely associated with the number of moves the family had made to a new home. Neighborhoods vary significantly in their safety, traffic congestion, availability of safe play areas, quality of schools, and access to other families with children, each of which can affect opportunities for children to create new relationships (Medrich, Roizen, Rubin, & Buckley, 1982). Parents are also gatekeepers who typically arrange, facilitate, and monitor their children's contact with others (O'Donnell & Steuve, 1983; Parke & Bhavnagri, 1989). Parents permit children to socialize with certain children and adults but not others. Parents schedule activities, provide transportation, and supervise offspring during social or recreational activities, especially when children are young (Ladd & Le Sieur, 1994). O'Donnell and Steuve (1983) found that lower-income and middle-income mothers differed significantly in the

access they provided their school-age children to community activities, with middle-income mothers participating extensively with their offspring in these programs (sometimes as volunteers and aides) in contrast with lower-income mothers, who instead permitted their children greater unscheduled freedom for “just being with friends.” Mothers in economic difficulty are less likely to have time off from work, access to transportation, or the resources necessary to invest themselves in their children's activities. Children's access to extended family members is similarly mediated by parental assistance, especially when kin are not local (Thompson, Scalora, Castrianno, & Limber, 1992). To be sure, children become increasingly competent at creating and maintaining their own social networks by choosing peers and activities, arranging their own transportation, and maintaining access through indirect means (such as technology) with increasing age. Parents remain important, however, as monitors and gatekeepers, and this role may be especially important for children and families at risk (Dishion & McMahon, 1998). With respect to developmental psychopathology, of course, the parental gatekeeping role means that access to extrafamilial sources of social or therapeutic support may be contingent on the parent's support for this access. The social networks of children and their parents overlap in other important ways. Parental social networks have significant consequences for children because they influence parents' well-being, provide opportunities for new experiences and relationships, and sometimes directly socialize and support parenting competence through modeling and mentoring (Cochran, 1990; Cochran & Niegro, 1995). Thus, when mothers receive social support from people in their social networks, it is associated with many positive outcomes for children, potentially including more secure attachment, more positive parent-child relationships, and improved peer relationships (e.g., Jacobson & Frye, 1991; Jennings, Stagg, & Connors, 1991; Melson, Ladd, & Hsu, 1993; see Parke & Buriel, 2006, for a review). These child outcomes arise, in part, from the benefits of social support for the quality of parenting, including mentoring and practical assistance (e.g., respite care). At times, moreover, parents' network associates can also be role models and natural mentors for children. Of course, parents' social networks can undermine rather than improve parenting practices, especially when extended family members, coworkers, neighbors, or others create stress rather than enhance well-being or provide reinforcement for poor parenting practices (Ceballo & McLoyd, 2002). As another illustration, then, of how relationships affect the functioning of other relationships, the social support or conflict that parents experience in their own social networks has important implications for the quality of their parent-child relationships and, in turn, children's development. Children's and parents' social networks overlap also because they are constructed in the same neighborhood and community context and are thus affected by comparable resources, risks, and stresses. As noted earlier, neighborhoods vary in the resources they provide for families with children (e.g., safe play spaces, recreational activities); equally importantly, they also vary in the human capital on which families can rely. When neighborhoods in economically disadvantaged communities are high in these human resources, children benefit significantly (Odgers et al., 2009). In most such neighborhoods, however, limited economic resources are combined with social impoverishment to undermine support resources for family members (Garbarino & Kostelny, 1992). In a classic study, Garbarino and Sherman (1980a, 1980b)

compared neighborhoods with higher-than-expected child maltreatment rates (based on sociodemographic predictors) with neighborhoods that had lower-than-expected maltreatment rates. Sampling informants in each neighborhood, they found that mothers in higher-risk neighborhoods reported receiving less assistance from neighbors, finding fewer options for child care, and generally perceiving the neighborhood as a poorer place for raising children. Risk to children was greater in neighborhoods perceived as providing fewer human resources for the families living in them. Children can also be affected in other ways by troubled neighborhoods. Lynch and Cicchetti (2002) reported, for example, that children who reported that they had been exposed to high levels of community violence also reported feeling less secure with their mothers. Understanding the functioning and developmental influence of children's social networks requires, therefore, perceiving them within the broader social ecology and the social networks of other family members. A study of Australian families by Homel, Burns, and Goodnow (1987) illustrates this. These researchers reported that the socioemotional adjustment of school-age offspring was predicted by the breadth and reliability of parental support networks, as indexed by the number of “dependable friends” that parents could list, and the parents' affiliation with voluntary organizations. However, the family's “neighborhood risk level” (indexed by the neighborhood's socioeconomic status, delinquency and school truancy rates, and related variables) also predicted child adjustment in the opposite direction, indicating that community-level as well as proximate features of the family's social ecology were important. Child well-being was affected both by the availability of immediately supportive partners to parents and by the broader human resources of the neighborhood. This conclusion is affirmed by the findings of another Australian study by Cotterell (1986), who found that the child-rearing attitudes and behavior of parents were strongly predicted by an interaction between (a) the father's absence from the home (owing to job demands), (b) the mother's “informational support” from neighbors about raising children, and (c) characteristics of the community (such as the stability of the population). Support to the mother was the most significant of these predictors, but all were important to parental conduct. It is important to remember, however, that children have their own neighborhood networks that may be relatively independent of the family. This network independence is especially likely when they reach adolescence, and those networks can buffer other characteristics of the neighborhood ecology if they function supportively. In one illustration, the internalizing symptomatology of a sample of urban African American adolescents was studied in relation to characteristics of their neighborhoods. In general, higher neighborhood poverty and unemployment rates predicted greater internalizing symptoms, and these associations were mediated by lower perceptions of neighborhood cohesion and perceived social support. However, when neighborhoods were residentially stable and had higher concentrations of African Americans, adolescents reported greater social support and higher perceptions of neighborhood cohesion, and these network characteristics were associated with fewer internalizing symptoms (Hurd, Stoddard, & Zimmerman, 2013).

Online Social Support

We close this section with the question with which it opened: Can the availability of social media, online social networks, the Twitterverse, blog posts, even virtual gaming and a growing range of other technological advances offer avenues of social support to youth and adults? In light of the foregoing discussion, it is now possible to identify more clearly some of the unique characteristics of social support through social media. Social media offers people opportunities for social contact that are convenient, self-determined, and easy to initiate. Requiring only a computer with an Internet connection, people can interact with individuals throughout the world with diverse backgrounds, perspectives, and experiences who would be inaccessible through other means, and they can also interact online with individuals whom they see on a regular basis. People can be selective in choosing these online contacts based on the websites, media, or privacy settings they use. Moreover, people can also be selective in the information they divulge about themselves, providing representations of themselves and their experiences that may be genuine, partial, or even fabricated, depending on their purposes in doing so. Perhaps for these reasons, social media appears to be a desirable avenue for obtaining emotional support from others, such as through text messaging, Facebook posts, or tweets. The individuals selected for contact can respond with concern and support to one's descriptions of their personal experiences. Although other forms of material or protective social support may be less possible, emotional support may be an important benefit derived from online media. Emotional affirmation is especially possible because some of the obstacles to social support posed by relationships discussed earlier, such as the stresses imposed by privacy intrusions or personal demands, are less likely when one can choose one's online contacts. Furthermore, it is possible that online communication could have other benefits for psychological well-being, such as obtaining information and guidance that socializes conduct in healthy ways. The explosive recent growth of social media and its use by older children, youth, and adults has provoked research attention by social and developmental psychologists. Research in this field is still in its infancy, but there are several lessons relevant to social support and developmental psychopathology, with a considerable research agenda remaining. There are at least two uses of online communication for social interaction. The first consists of online contact with individuals with whom one is in direct contact. In such cases, online media —such as instant messaging and texting, e-mail, Facebook, Twitter, and other kinds of social networking media (e.g., Spotify, Pinterest)—provide a means of ongoing digital communication or, in some cases, enables continuing communication with those who were previously in one's social network, such as friends who moved from the area. Despite early concerns that participation in online social networking would be problematic for youth, precipitating a decline in healthy personal relationships and increasing risk for loneliness and depression, the most recent research indicates that adolescents who are more engaged in online communication experience higher relationship satisfaction with people they are in direct contact with and enhanced psychological well-being (see Valkenburg & Peter, 2009b, for a review). The same may also be true of adults (Bessiére, Kiesler, Kraut, & Boneva, 2008). One explanation for these findings is that online communication fosters enhanced intimacy and reciprocity in self-disclosure, as people are more comfortable sharing sensitive personal

information with their friends in an online context, and this enhanced intimacy improves relationship quality with these friends (e.g., Valkenburg & Peter, 2009a; see also Jiang, Bazarova, & Hancock, 2011). Thus, online contact with friends strengthens relationship support. This process appears to particularly benefit socially anxious adolescents who use online communication to enhance relationships with friends, in part because of reduced social cues triggering anxiety and the opportunity to carefully craft communications (Valkenburg & Peter, 2007a, 2007b). For these youth especially, the ability to deepen existing friendships online may provide an important buffer against psychological difficulties exacerbated by challenges in direct peer relationships. Perhaps unfortunately, socially anxious youth are also somewhat less likely than their peers to engage in online communication with friends (Valkenburg & Peter, 2007b). A second use of online communication for social interaction is to initiate and maintain interaction with individuals who are known only online – such as through chat rooms, virtual gaming, blog posts, and other social media—although most adolescents and adults primarily use online tools to maintain and enhance existing personal (offline) friendships. Research findings are more mixed regarding the consequences of the Internet for relationship initiation and communication with unfamiliar peers. Some studies suggest that young people and adults who use the Internet primarily to communicate with contacts who are known only online may have reduced psychological well-being (Bessière et al., 2008; Valkenburg & Peter, 2007b). But there is also evidence that communication with online-only contacts can alleviate psychological distress. In one study, for example, online communication with an unfamiliar peer had a positive effect on restoring adolescents' self-esteem and perceived relational value, and reducing negative affect, following an episode of social exclusion (Gross, 2009). Additionally, unique processes may operate for individuals who are members of marginalized or stigmatized groups. For these people who have limited access to social support in their communities—and who have higher rates of depression and suicidal ideation—access to a larger virtual community can serve important support functions, reduce isolation, and aid in identity development (e.g., Gray, 2009; McKenna & Bargh, 1998). More broadly, the correlates of Internet use to form new relationships and connect with unfamiliar people may depend on the level of existing social support: People who have smaller social networks and limited social support available to them may benefit especially from social connections online (Bessière et al., 2008). There are also important developmental considerations in the function of online social networks as relevant to psychological well-being. Very little is known, for example, about the use of online resources for social support in preadolescence. In later adolescence and adulthood, however, social network sites become an aid in maintaining supportive connections to physically distant close friends, with positive implications for emotional adjustment (Ranney & Troop-Gordon, 2012). At the same time, social networking allows individuals to expand their network of new contacts easily. Greater use of social networking for this purpose appears to promote bridging social capital: the sense of being connected with and able to effectively elicit resources from a broad community (Ellison, Steinfield, & Lampe, 2007). Although this function of social networking diverges from the more intimate functions of

enhancing and deepening relationships, a sense of connectedness and access to support within a new community (e.g., during the transition to college) may still serve an important protective function, particularly during transitional developmental periods. Still, many questions remain regarding the nature of social support in online relationships and the connections between online support and developmental psychopathology. For example, although most people use online communication as an extension of offline friendships, it is important to better understand whether unique relationship processes relevant to social support occur in online communications between friends. It is known that disclosure is often more intimate in direct and online communication to support relationship closeness, but it is not known whether the negative effects of co-rumination could be exacerbated in an online context or whether friends are willing to address dysfunctional or inappropriate conduct online (e.g., Rose, Carlson, & Waller, 2007). Further, when youth exhibit signs of psychological distress in an online context, do those signals elicit support from others, and if so, what form does support take, and how effective is it? An intriguing recent study with young college students suggests that individuals signaling depressive symptoms online would still prefer that friends and acquaintances express their concerns in person, perhaps underscoring that when concerns are deeply personal and disclosure is risky, such communications should not take place online (Whitehill, Brockman, & Moreno, 2013). However, this study asked only how individuals might prefer someone to contact them rather than how friends and acquaintances actually respond. Finally, research on online relationships has primarily addressed informal connections rather than formal social support programs that might be created for people facing mental health challenges. This avenue merits further exploration, as some formal online support programs for individuals dealing with specific adjustment issues demonstrate benefits for participants. For example, Dunham and colleagues (1998) reported that adolescent mothers of infants who were given access to a computer-mediated social support network provided each other with positive emotional, informational, and tangible support. Mothers developed close personal relationships with each other, and active participation was related to a decrease in parenting stress over the 6-month study period. Given adolescents' comfort with online forums, the relative ease of establishing an online environment in which adolescents feel comfortable discussing personal vulnerabilities, and the decreased personal and social risk in such an environment, online social support networks or programs may hold promise for adolescents (see Caplan & Turner, 2007). However, much work remains to determine the utility of online connections—either informally or as part of formal social support systems—to buffer or alleviate distress for people at risk or experiencing psychological difficulty (see Hazzard, Celanno, Collins, & Markov, 2002, for one example for seriously ill children). Taken together, a considerable research agenda remains for better understanding the potential benefits and risks of social media and other online resources to youth and adults. But research evidence thus far suggests that initial concerns about its problematic consequences have not been confirmed, while some of the benefits of more intimate disclosure online may be meaningful and significant. Despite widely publicized accounts of the devastating consequences of online bullying by young adolescents, for example, the supportive and

potentially therapeutic uses of online resources merit further examination.

Giving and Receiving Social Support in a Cultural Context Social support is deeply integrated into cultural values and practices concerning individuals and their social connections. It is associated with cultural beliefs about the nature of the person and his or her social obligations, equity and reciprocity in social relationships, and family and parental responsibilities. This discussion of social support in cultural context extends the relational focus of the preceding section by first considering the challenges of giving and receiving support as it is experienced in mainstream Western culture. The focus is then widened to consider these issues from an expanded cultural lens, and the implications of this analysis for social support and developmental psychopathology are then evaluated.

Recipient and Helper Reactions to Assistance Receiving assistance from another evokes surprisingly mixed reactions from recipients in the United States, which is where most research on social support is conducted. In addition to the feelings of pleasure and gratitude that helping naturally inspires, recipients may also experience various negative feelings (J. D. Fisher et al., 1982; Shumaker & Brownell, 1984). Receiving assistance can be humiliating and stigmatizing, especially when the need for assistance derives from inadequacies in the recipient (such as poor parenting, substance abuse, or inadequate personal or financial management) rather than from broader, impersonal circumstances (such as an economic recession or a natural disaster) (Heller & Rook, 2001). Receiving help can also create feelings of failure, indebtedness, and inferiority, especially when assistance cannot be repaid, because of cultural norms of equity and reciprocity (Greenberg & Westcott, 1983). If assistance cannot be reciprocated or compensated, the recipient may also experience feelings of vulnerability or dependency because obtaining assistance from another violates Western norms of self-reliance and autonomy. There can also be sensitivity to privacy violations if helpers become intimately acquainted with aspects of the recipient's life that are not normally disclosed to others. As a consequence of these reactions, recipients may rather paradoxically begin to resent the assistance they receive and the person providing it. This resentment is especially likely when assistance is received from voluntary benefactors (whom one cannot give back to or otherwise compensate, enhancing the violation of equity and reciprocity norms) or strangers (with whom one does not share an ongoing relationship of mutual aid) and when the helper and the recipient are from similar backgrounds and circumstances (enhancing the inequity of the helping relationship). When recipients experience assistance as humiliating, demeaning, or intrusive for these reasons, they are less likely to seek help in the future and are more likely to abridge or terminate a helping relationship if they are capable of doing so. These reactions can explain why recipients of assistance, to the surprise of their benefactors, may be ungrateful, may fail to become engaged in the helping relationship, are often inexplicably absent from scheduled

meetings, do not return phone calls, and progressively make the relationship unworkable or unsatisfactory. This analysis has surprising implications for providing social support to at-risk individuals or families. It suggests that assistance is more easily accepted when recipients have opportunities to reciprocate or repay the aid they receive, perhaps in service to other individuals. It suggests that support is more readily received in circumstances that minimize the potential for humiliation or stigmatization, such as when support services are broadly available or universal (rather than specifically targeted to those in greatest need) and accessed in everyday settings (e.g., at home rather than at an agency office). This analysis suggests also that social support is more easily received when the recipient and the helper agree about the need for assistance and the reasons for the need. By contrast, assistance from another may be resented when the recipient perceives that it derives from unshared judgments of the recipient's inadequacy or incompetence. Provider efforts to preserve the dignity and the privacy of recipients are thus important. Interestingly, although negative recipient reactions to assistance are a risk when adults are concerned, they are much more rarely observed in children because of the universal acceptance of children's dependent status and their need for care. Children expect and readily accept the social support they receive from adults most of this time, and this is one reason why feelings of gratitude, as well as resentment, are not as characteristic of children's responses to the solicitude of their parents, teachers, or other adults. Children recognize that they need assistance and that adults are expected to provide it. To be sure, as children become more competent with increasing age, adult help that they regard as unwarranted can be threatening to self-esteem and perceptions of personal competence. Mothers often help their children with homework, for example, but with increasing age, children may begin to perceive that their mothers' helping, monitoring, and praising reflect mothers' beliefs about their children's lack of competence to do the work on their own (see Pomerantz, Moorman, & Litwack, 2007). Because children depend on adults for assistance of many kinds, especially when they are young, however, these resistant responses to help are fairly rare. The negative responses of adults to social support often interact with other recipient characteristics that can make it difficult for them to accept assistance, such as problems with substance abuse or affective psychopathology, personal disorganization, or the emotional effects of stressful circumstances (Heller & Swindle, 1983; Shinn et al., 1984). Taken together, these characteristics not only make it difficult for adults to maintain supportive helping relationships with other adults but also place considerable strain on help providers, especially in informal social networks. Offering assistance to troubled individuals can be draining and demoralizing (Collins & Pancoast, 1976; Shumaker & Brownell, 1984). Recipients have needs, but they may also be demanding and critical, and providing help in relationships of oneway assistance can be unrewarding and exhausting because support is not reciprocated. Moreover, the relationship between support providers and recipients can be difficult because each may have different goals, with recipients seeking noncritical emotional affirmation and providers striving for changes in recipient behavior and attitudes. They may differ in their views of the recipient's problems and the best solutions to them. Crises may force support

providers to focus on immediate needs (urged to do so by recipients) and neglect attention to long-term strategies for building healthy practices. For these reasons, it is common for providers and recipients each to feel frustrated by their relationship and sometimes to experience conflict. Informal social support can break down, and social support interventions enlisting natural helpers must address the frequent turnover and burnout of their staff. These vulnerabilities derive from the challenges presented by recipient responses to the help they are offered and the challenges of providing assistance to them in relationships of one-way aid.

Cultural Considerations in Giving and Receiving Social Support Although these challenges are, to some extent, intrinsic to providing and accepting social support, especially in informal social networks, cultural processes moderate them considerably (Dilworth-Anderson & Marshal, 1996; Jacobson, 1987). Understanding these processes is important for better conceptualizing the associations of social support with psychological well-being and developmental psychopathology, as well as for the design of effective interventions to enlist social support for children and families at risk. A well-known dimension by which concepts of the self in relation to others vary interculturally is that of individualism and collectivism (Triandis, 1989) or independence and interdependence (Markus & Kitayama, 1991). In cultures with an interdependent view of self, there is greater emphasis on connectedness with others and on deriving important features of identity and esteem from those associations. In cultures with a more independent view of self, there is a greater emphasis on the autonomy of personal thoughts and feelings, self-reliance, and privacy. Critics of this dual-cultures approach point out the heterogeneity of cultural views of the self within each orientation, as well as the fact that some societies making the transition to increased urbanization and industrialization integrate elements of each orientation (Kagitcibasi, 2005). In any case, cultural views of the self are developed quite early and influence how children perceive themselves and their relationships from early childhood (e.g., Greenfield, Keller, Fuligni, & Maynard, 2003; Keller, 2007; Rogoff, 2003). By later childhood and adolescence, youth with backgrounds from interdependent cultures (such as Hispanic and Asian societies) acknowledge the expectation that they will assist and support family members more than do adolescents from European backgrounds (Fuligni & Pedersen, 2002; Fuligni, Tseng, & Lam, 1999), and one study of Chinese American teenagers reported that such intergenerational expectations had neither positive nor negative consequences for psychological well-being (Fuligni, Yip, & Tseng, 2002). But the association between cultural values and family support is complex. A cultural emphasis on interdependence may facilitate help-giving and help-receiving through normative practices, but cultural values may also make receiving help more difficult in many circumstances. One illustration is a study of older Japanese American adults living in New York City for whom norms of reciprocating support made receiving assistance difficult. Adults who held strong reciprocity norms and who received material support from their families were more depressed and were less satisfied with their lives than those who did not embrace strong reciprocity norms (Nemoto, 1998). Thus, in a cultural context that emphasizes interdependence, negative recipient reactions to receiving assistance nevertheless emerged owing to the reciprocity

norms of that culture. Cultural values and practices are related to a number of features of social networks and social support. Specifically, there is significant intercultural variability within the United States in (a) the nature and functioning of informal social networks, (b) the association between social support and psychological well-being, and (c) attitudes toward receiving assistance from formal helpers. Each of these sources of variability is relevant to designing culturally competent interventions involving social support and linking formal and informal sources of support to troubled children and their families, as described below. Although the components of social networks are similar for families in different ethnic and cultural groups (such as including immediate and extended family, friends, neighbors), the relative importance of each of these network members for social support can vary. MacPhee, Fritz, and Miller-Heyl (1996) compared lower-income Native American, Hispanic, and European American parents living in the United States on their self-reported sources of support. They found that Native Americans reported more interconnected social systems, more frequent contact with extended kin (but not friends), more members who knew one another, and greater closeness with members of their support networks. Hispanic parents reported having the largest social networks and, although these networks were close-knit, Hispanics were in general most likely to rely primarily on kinship networks for emotional support. While they reported the lowest proportion of network members who could offer emotional support, they also reported the highest proportion of network members who could provide instrumental support (e.g., material assistance). European American parents had more diffuse social networks but also reported having a higher proportion of members available for instrumental support. Unlike parents in the other two groups, they reported that friends were the primary providers of emotional support, not family members. Children also exhibit intercultural variability in the network members on whom they rely for support. For Latino and African American children, for example, support may derive from extended family or in broader fictive kin networks (e.g., Rhodes, Ebert, & Fischer, 1992; Sanchez & Reyes, 1999;). In African American families, children and youth report receiving greater support from their extended families than do children from European families (Cauce, Felner, & Primavera, 1982; Taylor, Casten, & Flickinger, 1993). Likewise, Hispanic values of familialism, involving strong feelings of support and reciprocity with family members, expand the social support networks of children to include adults beyond the immediate family unit (Sagobal, Marin, Otero-Sabogal, Van Oss Marin, & Perez Stable, 1987). These intercultural differences are important for understanding the network members who are likely to provide the most helpful forms of informal social support to children and families in need, as well as the avenues by which such support can be offered. Although social support comes from potentially diverse sources and is experienced within the context of cultural values, there is good evidence that despite this, social support contributes to psychological well-being for different cultural groups within the United States. Coatsworth and colleagues (2002) examined family, school, and friend support in relation to externalizing and internalizing behavior in Hispanic girls in middle school, for example, and found that

controlling for age, socioeconomic status, and years in the United States, youth reports of greater perceived family support and teacher support (but not friend support) predicted fewer externalizing problems, while greater perceived family support and friend support (but not teacher support) predicted fewer internalizing problems. Support within the family was the strongest buffer of externalizing and internalizing symptomatology. Likewise, in a short-term longitudinal study of African American male adolescents, Zimmerman Ramirez-Valles, Zapert, and Maton (2000) found that although support from friends did not predict later outcomes, support from parents predicted diminished depression and anxiety. Rodriguez, Bingham Mira, Myers, Morris, and Cardoza (2003) examined perceptions of family and friend support in relation to stress and psychological adjustment in Latino college students. They found that higher support from family and friends predicted increased psychological well-being, although only friend support was a unique predictor of lower psychological distress. However, cultural values may significantly mediate whether members of different ethnic and cultural minorities access formal—rather than informal—supports when facing psychological distress. This difference between informal and formal supports can occur for various reasons, including lack of awareness of formal services, distrust of providers (or providers who cannot speak their language), cultural beliefs that assistance from nonfamilial helpers is unnecessary or inappropriate, resistance to formal helpers from within the family or cultural group, service delivery practices that are culturally uninformed, or for other reasons. In the study of lowerincome families earlier described, MacPhee and colleagues (1996) found that European American parents were significantly more likely to have sought professional therapy than were Native American or Hispanic parents, and they also tended more to seek professional help with parenting issues. Findings such as these underscore the need for cultural awareness in designing interventions to enhance social support to children and families of culturally diverse groups. A culturally competent service delivery system will (a) identify groups that are underserved and seek to reduce cultural barriers that may interfere with service delivery by understanding their characteristics, resources, and needs; (b) orient program planning, staff training, and community involvement to ensure that the development, implementation, and evaluation of services are respectful of the values and practices of recipient families; (c) evaluate assessment and outcome procedures and instruments to ensure their appropriateness and validity for the children and families who are served; (d) build cross-cultural communication skills within program staff, including the appropriate use of interpreters and an ethnographic understanding of communication approaches within cultural groups; and (e) seek to develop an appreciation of cultural diversity as a facilitator rather than impediment to service delivery. These practices are especially important in services that seek to strengthen the benefits of social support interventions by linking formal support to informal support networks, especially in light of how “outside” helpers can be regarded with distrust or resentment by members of closely knit families or communities. In the end, cultural beliefs and practices are among the most significant personal characteristics mediating the needs of potential recipients and the providers of social support.

Interim Conclusions Before turning to research focused on social support and developmental psychopathology, there is value in summarizing the implications of the foregoing analysis of social support in the lives of children, parents, and families. Doing so provides a set of conclusions that suggest the kinds of socially supportive interventions that are likely to prove most effective in helping children and their families. Several conclusions and their implications seem warranted: Relationships influence other relationships. Parents mediate children's access to other people outside the family and the support they can provide. The quality of parent-child relationships is affected by the support or conflict that parents experience within their own social networks. The ability of a formal helper, such as a home visitor, to engage parents and children may be affected by attitudes of extended kin. These conclusions underscore the interconnected nature of the social networks of family members and how relational influences on one person have broader scope for how that person engages others. Insofar as social support is given and received in the context of relationships, these observations suggest that supportive relationships beget other supportive relationships—with the opposite result when relationships are stressful and taxing. With respect to intervention, this conclusion suggests that multigenerational interventions are important to strengthening social support to children, as well as efforts to improve parental social networks. Improving parental well-being, such as through the assistance of a grandparent, a friendship network, or a professional helper, can contribute to improving the quality of parenting and, through this, children's well-being. Because child psychopathology is typically associated with broader family processes, intervening to improve the quality of multigenerational relationships within the family, and the quality of the social networks that each family member draws on outside the family, improves the chances for healthy development for children within that family. Coordinating informal and formal sources of social support is important. Formal and informal helpers have, as we have seen, unique strengths and weaknesses in the provision of social support (Gottlieb, 1983, 2000). Members of informal social networks share the values of recipients, and their accessibility, mutuality of assistance, trust, and multifaceted avenues of potential assistance make them influential agents in everyday life. But because natural helpers may share the circumstances of their recipients and lack extensive training, they may be ineffective and are subject to exhaustion and withdrawal as helpers. Formal sources of support, by contrast, have professional training and a formal role definition that contributes to their impact and the reliability of their assistance. But because they are less well integrated into the lives of recipients and do not share their values, experiences, and background, support can be stigmatizing, and formal helpers may have difficulty engaging the cooperation of family members. Because of these distinct characteristics, integrating the efforts of formal helpers with those of informal helpers in recipients' natural social networks may offer the best opportunities for creating enduring preventive or therapeutic benefits (Froland, Pancoast, Chapman, & Kimboko, 1981; J. L. Miller & Whittaker, 1988). Formal helpers working with members of

informal social networks can support natural helpers in their efforts while ensuring the skill and expertise that formal helpers can provide. The integrated efforts begin with the mutual recognition that formal and informal helpers have unique contributions to offer needy recipients, and this mutual recognition can occur in many ways. Formal and informal assistance is harmonized, for example, when an early childhood mental health consultant works with parents or teachers at a local school or child care program, a perinatal home visitor encourages the involvement of extended kin during home visits, or a group therapy program for adolescents has connections to the school or to the peer group. In child welfare, children and/or family members are often encouraged to enlist someone from their natural social networks to advocate for them at social service meetings and help them follow through on the goals identified at these meetings. The effective coordination of formal and informal support networks is not easy, however, because of the differences in background, values, goals, and definition of the problem that may provoke mutual distrust between formal and informal helpers. All too commonly, for example, extended family members or neighbors reinforce a parent's skepticism regarding the potential helpfulness of a counselor or paraprofessional home visitor. Sometimes social workers undermine informal helpers by criticizing them to the recipient. But the integration of formal and informal helping is important to promoting the engagement of recipients in social support interventions and to providing a foundation for continuing assistance. Many well-meaning social support interventions fail because they do not sufficiently incorporate the natural helping networks of family members, or rely on them exclusively, resulting in assistance that is limited in time, scope, and impact. Parents are central to children's environment of relationships, but they are not solely influential. The first part of this conclusion affirms the importance of parent-child relationships to children's developmental competence, emotional well-being, and capacities for receiving social support. It also underscores the vulnerability of children, especially when they are young, to parental dysfunction. Research on children in homes characterized by marital conflict shows, for example, that children become emotionally enmeshed in their parents' conflict, contributing to risk for their own emotional psychopathology. They monitor parental moods to anticipate conflict before it occurs, and they intervene in parental arguments as they are trying to manage their own emotions (Davies, Harold, Goeke-Morey, & Cummings, 2002; Davies & Woitach, 2008). The emotional dysfunction in a child created by a parent who has become a source of threat or stress is illustrated in research on the effects of parental depression, domestic violence, and, of course, child maltreatment. In the attachment literature, it is often observed in young children developing insecure disorganized attachments, manifested behaviorally in the approach-avoidance conflict of a child who needs but cannot seek assistance from the parent (Lyons-Ruth & Jacobvitz, 2008). In these and other circumstances, the social support that children need from within the family is inaccessible. Moreover, social support from outside the family also may be inaccessible because of how parents function as gatekeepers to children's extrafamilial social contacts. Parents may limit children's access to alternative support agents because of their own emotional problems or to conceal their

disapproved conduct from detection. The recognition that parents are central to children's emotional well-being, therefore, underscores not only the importance of the support that parents ordinarily provide but also the vulnerability of children to parental dysfunction. The second part of this conclusion, however, underscores that others can have important influence in assisting children even when they cannot ameliorate negative family experiences. Indeed, some research suggests that support from only one or two people outside the family can be instrumental in helping children in stressful family circumstances cope more effectively (Beeman, 2001). As this review has shown, those extrafamilial agents may be found in early child care settings, primary grade classrooms, and community and recreational activities; for older children, they may be found in the peer, after-school program, and workplace environments. When children are young, extrafamilial support agents can provide them with a safe haven of security and predictability that can support coping even when the family environment remains difficult. At best, such agents can help bridge the provision of other resources to the family (Small, 2009). When children are older, programs designed explicitly to provide support, such as youth mentoring programs, have the potential of providing considerable assistance (Hirsch, 2005). Thus, parents are central but other agents of social support can also be influential in helping children at risk. Representations of support from others are important to psychological well-being. The importance of perceived support to psychological functioning suggests that expectations that support is available are at least as important as the actual experience of receiving assistance from network associates. Indeed, perceived support seems to be more important than enacted support in many circumstances. In light of the modest association between perceived and received support, this conclusion warrants further research into the influences that cause individuals to expect that others will provide assistance when needed, even if they have not regularly obtained such aid in the recent past. How affective psychopathology influences individuals' perceptions of support, which may be independent of the actual assistance they receive from formal and informal helpers, also warrants research attention. At the very least, this conclusion means that interventions to strengthen social support to increase psychological well-being must attend as much to changes in recipients' expectations for future support as to changes in the actual support they receive because the benefits of intervention may be limited unless support expectations are altered. In this respect, it is important to recognize that expectations of support from significant people have a stress-buffering consequence. One reason why attention to perceived support is so important is that it requires addressing why recipients may not perceive support in contexts when it is provided. It is likely that natural helpers and formal helpers who focus attention on the recipient's dysfunctional behavior and try to help change it may be perceived as unsupportive, even if their efforts are necessary and potentially effective. The research on negative recipient reactions to assistance suggests also that how support is provided may be significant for shaping recipients' reactions to the assistance they receive and their expectations for future support. When support provision is normalized for the recipient's neighborhood or community, when it is provided in contexts that avoid stigma (e.g., at home or church rather than a

clinic), when it is broadly available rather than targeted, when recipients can repay the help they are given, and when helpers and recipients jointly agree on the need for assistance and the goals it will accomplish, it is more likely that received support will be perceived as such. Furthermore, when formal helping professionals take the time to establish relationships of trust with recipients and respect their perspectives and values, their assistance is more likely to be regarded as supportive. Understanding how to enhance perceptions of supportiveness in the context of social support interventions is an important future research task. Little is known, furthermore, about developmental changes in expectations of social support. Research reviewed above suggests that securely attached children perceive greater support from those around them, but it is unclear whether such perceptions derive from greater amounts of actual assistance (deriving perhaps from the capacity of secure children to elicit support from others) or whether a secure attachment itself contributes to more positive expectations. Very little is known about how other child characteristics, including temperamental individuality, contribute to perceptions of social support, or about developmental changes in these expectations. It is also crucial to understand how children's actual experiences in the family, such as marital conflict, affect children's generalized expectations of social support and how they are affected by potentially stress-buffering relational influences, such as through interactions with other adults. There is potentially a very rich research agenda here with important practical applications and clinical implications. Social support efforts will be more effective when there is clarity concerning the goals they are intended to achieve. Because social support is multifaceted and recipients have different constellations of challenges, a one-size-fits-all strategy is inappropriate to social support interventions. Rather, interventions should be preceded by a careful analysis of the needs of target individuals or families, the capabilities of informal or formal helpers enlisted for them, the specific purposes or goals of supportive interventions, and the outcomes that can reasonably be expected to result (Gottlieb, 2000; Heller & Rook, 2001). Such an analysis is important is because there are inevitable trade-offs between different strategies and goals that might be envisioned. Enlisting formal helpers is typically more costly than using volunteers and/or informal network members, but it may be necessary and cost-effective when the recipient's problems are so serious that they are beyond the capabilities of nonprofessionals to address effectively. Universal or broad-scale support programs can be useful in the context of expectations that recipients need information and advice but do not have serious problems, but determining what outcomes might reasonably be anticipated from such programs can be difficult. At times, other resources may have to be enlisted in addition to supportive interventions, such as social skills training (Gaudin, Wodarski, Arkinson, & Avery, 1990–1991) or public assistance (Pelton, 1989, 1994) in order to address problems that contribute to the recipient's difficulties. Taken together, therefore, a generalized initiative to broadly provide support with the hope that it will improve the well-being of children or their parents is less likely to be successful than one that is based on a careful analysis of the child's and family's needs, who can provide

assistance, and the reasons for intervening. Stress and social support are complexly interwoven. The prevailing approaches to understanding the association between social support and psychological well-being are to regard social support as either stress preventive—through the incorporation of healthy practices into the recipient's experience—or as a stress buffer, supporting more effective coping by the recipient. But this analysis has shown that the association between social support and stress is much more complex. Stress may undermine the capacity of recipients to solicit or receive support and may increase their negative reactions to receiving help. As it is shared among neighborhood or community members, stress may also impair natural helping networks from providing mutual aid as individuals must cope with the demands of poverty, dangerous neighborhoods, or other difficulties in their own lives. The experience of receiving social support may itself be stressful, especially if it is stigmatizing or humiliating or contributes to feelings of failure or vulnerability. Recipients may also perceive helpers' interventions as unsupportive if it casts light on harmful practices that must change. Moreover, the process of providing social support may be stressful to helpers, especially if it becomes draining and demoralizing. In light of these considerations, there is value in distinguishing different sources of stress that social support interventions might address in potentially different ways. One source of stress is associated with the helper-recipient relationship, whether this is a formal professional association or an informal relationship between neighbors or friends. In each case, efforts to reduce the negative consequences of accepting aid may be necessary. Such efforts can include promoting a relational atmosphere of mutual respect and creating ways of normalizing support within the community in which the recipient lives. Another source of stress consists of neighborhood or community stressors, such as economic impoverishment and the drain of human capital resources, which contribute to more difficult lives. These stressors can have widespread influence within natural social networks and may be difficult to ameliorate, but they must be recognized and accommodated as the context in which support is provided. Finally, there are critical events, such as immediate crises, that a social support intervention can hope to change through effective intervention. By distinguishing different sources of stress, impediments to the efficacy of social support efforts can be recognized and, where possible, reduced to increase the effectiveness of interventions. Cultural awareness. Finally, cultural beliefs, values, and practices significantly moderate how social support is received, how potential sources of support are perceived (including attitudes toward formal as well as informal support agents), and the association between social support and psychological well-being. Developing culturally informed social support interventions requires recognizing the importance of these beliefs and practices, incorporating them into program goals and design, reducing cultural barriers, developing staff training that enhances cross-cultural communication, and valuing and respecting the beliefs and values of recipient families. Social support alone is unlikely to be an effective answer to the complex problems faced by

children with serious affective psychopathology or dysfunctional families. But social support is an important component of any multifaceted effort to prevent psychological difficulties in atrisk families and children or to provide therapeutic assistance when family problems have resulted in a child with clinical disorders. The challenge is to craft well-designed interventions that improve the support afforded by natural helpers and to enlist the assistance of formal helping professionals, based on a careful assessment of the needs of the child or family at risk (Thompson & Ontai, 2000). Accomplishing this requires a thoughtful appreciation of the nature of social support, as well as the specific role of social support in developmental psychopathology, to which we now turn.

Social Support and Developmental Psychopathology With these interim conclusions from research on social support in mind, we now turn to the research literature on developmental psychopathology. Recognizing the breadth of intervention efforts that incorporate social support, our purpose in this selective review is to focus on concepts and interventions that test and extend the conclusions described in the preceding section. To what extent, in other words, do existing preventive programs and clinical interventions draw on these conclusions, and in what ways could existing interventions be improved with greater attention to them? In this section, we discuss the relevance of social support to three phases in the course of psychological disorders. First, we consider how the absence of social support contributes to the development of psychological disorders. This is especially important to portrayals of social support as a preventive and buffering agent in stressful circumstances, and therefore we also consider the nature of preventive interventions to at-risk families lacking social support. Second, we examine the role of social support (and its absence) in the maintenance of psychopathology over time, underscoring the importance of social factors in the persistence of affective symptomatology. Finally, we consider social support and the treatment of psychopathology and the alternative avenues that exist for enhancing support as a therapeutic aid.

Social Support, Social Isolation, and the Origins of Developmental Psychopathology In light of the stress-preventive and stress-buffering functions of social support, it is reasonable to expect that children in families experiencing social isolation would be at greater risk of psychological difficulty than children in families experiencing considerable social support. Research on social isolation and developmental psychopathology, particularly studies of risk for child maltreatment, indicate the complexity of the concept of social isolation, however. Studies of efforts to address social isolation, particularly through home visitation, illustrate the challenges of engaging families in supportive social interventions to prevent the emergence of child and family problems. Social Isolation and Risk for Child Maltreatment

The social isolation of families who have abused or neglected their offspring, or who are at significant risk of doing so, is the most extensively studied condition in which the absence of social support is believed to contribute to risk for the development of child psychopathology. In this case, children's risk of clinical problems is believed to be mediated by their parents' lack of social support. Many researchers have concluded that parents who abuse or neglect their offspring lack significant social connections to others in the extended family, neighborhood, and broader community and to social agencies that can provide assistance (e.g., Daro, 1988; Garbarino & Sherman, 1980a; Polansky et al., 1981; Seagull, 1987, Gelles, & Steinmetz, 1980; see Thompson, 1995, for a review). As a consequence, their maltreatment of their offspring is likely to remain undetected, these parents have few interpersonal resources to which they can turn when they are stressed, and the ways that social connections with potential helpers can buffer the effects of stress, promote healthy behavior, and socialize positive parenting are strained. An extensive review of this research by Thompson (1995) yields, however, a fairly complex picture of the association between social isolation and child maltreatment. Two conclusions from his review are important to this discussion of social support and developmental psychopathology. First, in most studies, the social isolation distinguishing abusive or high-risk parents consists of their smaller social networks or their more limited social contacts with network members (i.e., limited social embeddedness). Parents at risk of child abuse know fewer people, and see them less often, compared to other parents in similar circumstances. But research findings are inconsistent about whether at-risk or abusive parents experience significant deficits in enacted support or have lower perceptions of expected support from their network associates. A study by Lovell and Hawkins (1988) is typical; in it, abusive mothers reported that very few of their network associates provided practical help with child care or parenting responsibilities, but mothers reported enjoying seeing nearly 80% of these companions very much and reported that they could “share their thoughts and feelings frequently” with nearly 50% of them. Like the fatally abusive mothers studied by Korbin (1989, 1991), in other words, parents often perceived support from network associates as satisfactory even though the social support mothers received did not significantly reduce abuse potential (cf. Freisthler et al.'s, 2014, “dark side of social support”). Thus, on the most important dimension of social support for psychological well-being—perceived support— there are often negligible differences between maltreating parents and those who are nonabusive, even though the social networks of maltreating parents are smaller and less supportive in other ways. Second, there are subgroups of maltreating parents who experience significant social isolation for specific reasons. Polansky and his colleagues (Polansky, Ammons, & Gaudin, 1985; Polansky et al., 1981; Polansky, Gaudin, Ammons, & David, 1985) have studied neglectful mothers, who report feeling greater loneliness and lack of neighborhood support compared with socioeconomically comparable nonneglectful mothers. They describe an “apathy-futility syndrome” consisting of a passive, withdrawn demeanor coupled with emotional “numbness,” limited competence, distrust of others, retreat from social contact, and verbal “inaccessibility” to others that makes them hard to reach socially (Gaudin & Polansky, 1986; Polansky &

Gaudin, 1983). In Polansky's view, the social isolation of these mothers derives from their general inability to develop and maintain supportive social ties owing to character disorders, deficient social skills, and difficulties in coping adaptively with life stress. In a sense, their neglect of offspring is part of a broader syndrome associated with their disorganization and lack of social and emotional competence. Garbarino has described, by contrast, a different kind of social impoverishment of families within neighborhoods that experience heightened rates of child maltreatment (Garbarino & Kostelny, 1992). As earlier noted, by comparing neighborhoods with higher-than-expected child maltreatment rates (based on sociodemographic predictors) with neighborhoods with lower-than-expected maltreatment rates, and using informants within each community, Garbarino sought to characterize the neighborhood conditions associated with child abuse and neglect (Garbarino & Sherman, 1980a, 1980b). He found that mothers of higher-risk neighborhoods reported receiving less assistance from neighbors and regarded the neighborhood as a poorer place for raising children, but on other assessments related to social support, such as perceptions of sources of potential assistance, mothers of higher-risk and lower-risk neighborhoods did not differ. By contrast with the characterological problems of the neglectful mothers studied by Polansky and colleagues, therefore, the mothers in the highrisk neighborhoods studied by Garbarino suffered from the diminished social resources of their neighborhoods and the diminished human capital of their communities. Studies like these indicate that social isolation is not a homogeneous phenomenon. Consequently the reasons for social isolation are diverse in families at risk for child maltreatment (Thompson, 1995). For some people, such as Polansky's neglectful mothers, isolation may derive from social marginality; for others, such as Garbarino's higher-risk neighborhood residents, it arises from the impoverishment of social capital in difficult neighborhoods. Some people may actively seek social isolation as a means of concealing abusive practices. For others, isolation may result from difficult circumstances that create feelings of humiliation and vulnerability and rob adults in the neighborhood of the time or energy required to maintain social networks. A significant proportion of high-risk families do not feel socially isolated at all; instead, they are satisfied with their social interactions with a small network of close associates who provide emotional support that does not seem to constrain abusive or neglectful practices. Social isolation, when it is apparent, can have diverse origins. The multifaceted causes of social isolation in families at risk for child maltreatment are important for at least two reasons. First, these studies indicate that social insularity may be a significant factor in the origins of child maltreatment for some families, especially when the causes of social isolation derive from psychological problems in parents, or difficult or dangerous neighborhood conditions, or active efforts to conceal abusive practices. In these circumstances, the absence of significant social connections increases risk for child maltreatment because there are few from outside the family who can provide assistance or monitor or socialize parental conduct, especially in the context of life stress. However, it is unwarranted to generalize this portrayal of abusive families. For many other families, social isolation does not appear to be etiologically relevant because abusive practices occur in the

context of active social networks from which parents derive emotional support, but which are benign influences on abusive practices. Social isolation is not necessarily implicated in child maltreatment—and perceived support is not necessarily a buffer against abusive parenting. Second, the diverse causes of social isolation are also relevant to intervention. Strategies for enhancing social support for at-risk families who are socially isolated must also be multifaceted. They may require, for example, social skills training (Gaudin et al., 1990–1991), improving recipient reactions to receiving assistance (Tracy, Whittaker, Boylan, Neitman, & Overstreet, 1995), incorporating new support agents into natural social networks, or other approaches depending on the causes of social insularity. When enlisting social support into clinical treatment, one size does not fit all. Risk for child maltreatment is not the only vulnerability for children living in socially isolated families because in these contexts, parenting quality is more likely to be poor, the family's use of community resources is more likely to be limited, and children's developmental outcomes are more likely to be problematic as the result of limited access to the mentoring and advice, information and material aid, developmental guidance and social monitoring that social networks can provide. This diversity of risks associated with family social isolation suggests that efforts to address it need to be clear about how social support is to be provided and the goals it is intended to achieve in order to be effective. This is well illustrated in the research on home visitation. Home Visitation as a Preventive Intervention During the past several decades, home visitation has become the most enthusiastically supported and widely recognized approach to providing social support for at-risk families. The fundamental strategy uniting diverse home visitation programs is the delivery of information, guidance, and emotional support to family members in their homes by a paraprofessional or volunteer home visitor. The purpose is to prevent the emergence of problems that may arise from parental stress, lack of knowledge or resources, or social isolation and thus to promote more healthy developmental outcomes beginning early in life. Home visits can begin prenatally, shortly after the child's birth, or later in childhood, and visits can extend in duration based on family need and engagement and program funding. Home visits are believed to provide an avenue for offering diverse forms of social support to parents and children, overcoming some of the barriers these families otherwise face to obtaining needed assistance (such as lack of transportation or health insurance), enabling monitoring of children's development, and establishing a relationship of trust with a home visitor who can provide individualized assistance and bridge connections to broader resources (Wasik & Bryant, 2000). Home visitation is the umbrella orientation of a number of programs established in communities throughout the country, most notably the Healthy Families America initiative, developed by the National Committee to Prevent Child Abuse, which has established a nationwide consortium of hundreds of home visitation programs serving families at risk (Daro, 2000; Daro & Harding, 1999). Home visitation programs are funded by direct legislative appropriation in many states, or by project grants from federal agencies, as central features of statewide efforts to strengthen child health and development, prevent child maltreatment,

improve parent-child relationships, or address other issues. Most notably, as part of the Patient Protection and Affordable Care Act (ACA), Congress in 2010 established a new Maternal, Infant, and Early Childhood Home Visiting Program and appropriated $1.5 billion over 5 years to develop and improve state-administered home visitation programs. Home visiting has considerable potential to support families with young children, especially those with limited social networks of support, but problems exist in the effective implementation of home visitation initiatives. This is one of the reasons for the mandate of the Obama administration, through ACA, that states focus on programs offering evidence-based documentation of effectiveness. A recent review of 35 different home visiting programs commissioned from Mathematica by the U.S. Department of Health and Human Services (DHHS), however, identified only 14 programs meeting the criteria for an evidence-based early childhood home visiting service delivery model (Avellar, Paulsell, Sama-Miller, & Del Grosso, 2013). Programs were classified in this manner because they had a high- or moderately high-quality research evaluation (e.g., randomized controlled trials, quasiexperimental designs, implementation studies) and demonstrated statistically significant positive effects in at least 2 of 8 targeted areas of impact, as well as sustained positive effects for at least 1 year following program enrollment. The impact areas identified in this report reflect the diversity of developmental outcomes that home visitation programs are intended to address: child health; child development and school readiness; family economic selfsufficiency; linkages and referrals to community programs; maternal health; positive parenting practices; reductions in child maltreatment; and reductions in juvenile delinquency, family violence, and crime. Across the evidence-based programs, the most frequent positive outcomes occurred in child development and school readiness, positive parenting practices, and child and maternal health. Positive outcomes were not limited to particular subgroups of families, and many were sustained for at least 1 year following program completion. In some cases, much longer-term impacts were identified. For example, the Nurse Family Partnership, which provides support to first-time expectant mothers beginning prenatally through the child's second birthday, had initial positive impacts on parental care and maternal behavior (e.g., subsequent pregnancies, education, and work). At a follow-up at age 15, children of program participants who had been low income and unmarried during program participation had fewer arrests and convictions, less substance use, and lower levels of “promiscuous” sexual activity compared with children of control group families from similar backgrounds (Olds, 2006). The Healthy Families America initiative described earlier also had positive outcomes in all assessment areas. Nonetheless, even the programs identified as most successful had relatively few effects relative to the range of targeted outcomes, and less frequently demonstrated positive impacts in areas such as family economic self-sufficiency, maternal and child health, linkages and referrals to community programs, and reductions in juvenile delinquency and family violence. (For many programs, these outcomes were not measured, however, because they were not part of the program's intended outcomes.) Several of these studies identified at least one unfavorable impact, although the nature of these impacts was not specified. Additionally, other programs reviewed did not meet criteria because they showed few positive effects or did not

sustain positive effects over time. The DHHS review did not identify the reasons for limited program success, but this has been the focus of other reviews of the home visitation research literature, including a 2004 metaanalysis of 60 home visitation programs by Sweet and Appelbaum (2004) and an influential assessment by Gomby, Culross, and Behrman (1999; see also Halpern, 2000). The reasons for limited program success identified in these reports include the inconsistent participation of recipients, the importance of supporting help providers, the need to develop community connections, and the problem of lack of clarity in program goals and expectations. The failure to fully engage families in the program and the high attrition rates of participants have been identified as significant challenges for virtually all home visitation programs. According to Gomby and colleagues (1999), between 10% and 25% of families invited to participate in home visitation programs decline and between 20% and 67% of the families who enroll fail to complete the program. Moreover, even when families enroll and remain in home visitation programs, they tend to receive only about half or fewer of the planned number of contacts with the home visitor (Gomby et al., 1999). Program engagement is particularly problematic for the highest-risk families, which may explain the more limited effects seen at higher levels of family risk. For example, Early Head Start evaluators reported that families with three or more identified sociodemographic risk factors were harder to engage and remained enrolled for shorter durations compared with families facing fewer risks (Administration for Children and Families, 2002). The reasons for problems in participant engagement are diverse, often related to residential relocation, busy or disorganized family schedules, and other characteristics of recipients. But lack of engagement may also be related to the mixed recipient reactions to obtaining assistance discussed earlier, especially if family members do not perceive that home visitation addresses their needs and concerns or if they feel embarrassed, indebted, or vulnerable because of the services they receive. A second problem in successfully implementing home visitation programs is the lack of training, supervision, and support for home visitors, which contributes to the high turnover rates that are observed in most home visitation programs and diminished program efficacy (General Accounting Office, 1990; Gomby et al., 1999). It is common for home visitors to report shorter visits than planned, broken appointments that are not rescheduled, and preoccupation with immediate family crises rather than the delivery of intended education or guidance during home visits. Home visitors are further challenged when working with culturally or linguistically diverse families, at-risk populations, or parents who suffer from depression, domestic violence, or substance abuse (Administration for Children and Families, 2002; Margie & Phillips, 1999). Moreover, several studies have found that how the intended curriculum of a home visitation intervention is implemented varies significantly depending on the values and orientation of the home visitor (Baker, Piotrkowski, & Brooks-Gunn, 1999; Wagner & Clayton, 1999). At times, in other words, what actually occurs during home visitations may be much different than what program designers had intended. The high turnover of home visitors further undermines the relationship between participants and the program, and this factor may contribute to the lack of family engagement. Turnover can be especially difficult for adults in the highest-risk families, who may have fewer alternative sources of support on

which to rely and who often have histories of abandonment and relational dysfunction (Goodvin, 2004, June]). These challenges to program implementation are directly related to the training, supervision, and support provided to home visitors, especially those who are volunteers or paraprofessionals. However, personnel, training, and supervision expenses account for most of the costs of a home visitation program, and thus poorly or inconsistently funded programs are likely to scrimp on these essential features of service delivery. This fact is especially unfortunate in light of findings that the professional identity of home visitors may substantively impact program results. For example, one evaluation of the Nurse Family Partnership— identified as a highly successful program for the breadth of positive impacts across target areas (Avellar et al., 2013)—included a comparison of nurses to paraprofessionals. The goal was to identify whether trained paraprofessionals could replicate the positive effects seen in earlier studies with nurse practitioners. However, results indicated that effects were not comparable: When paraprofessional home visitation produced any positive effects, they were approximately half the magnitude of those found with nurse home visitors (Olds, 2006). It is likely that in addition to their greater clinical training to support families dealing with challenges, nurses are also perceived by support recipients as having a greater degree of authority and useful information. But because of their professional training as formal helpers, nurse practitioners may not share the backgrounds, experiences, and often values of recipient families. Balancing the benefits of expertise with the value of shared orientation—which distinguishes formal from informal helpers—remains one of the challenges of designing effective home visitation programs. Another challenge to effective program implementation is the failure of many home visitation programs to explicitly establish the development of community supports as a central goal for recipient families. Illustrating this, only 4 of the 35 programs reviewed by Avellar and colleagues (2013) assessed whether program participation increased family linkages and referrals to other resources. This finding is unfortunate because the social support approach incorporated into home visitation recognizes that a home visitor cannot provide all that recipients need. Consequently one of the significant goals of intervention must be to help families forge associations within their communities to individuals and agencies that can offer longer-term support. Moreover, community connections and visibility can enhance the positive regard for a home visitation program within the neighborhood, which can also contribute to improving family engagement and strengthening the connections between family members and community services. Finally, the reviews by Gomby and colleagues (1999) and Sweet and Appelbaum (2004) each identified program goals as problematic because they were unclear, unduly ambitious, and/or not carefully translated into intervention strategies. Gomby and colleagues emphasized, in particular, the need for a renewed appreciation that home visitation programs must have modest expectations for what social support alone can accomplish for needy families. Thus, home visitation efforts should be combined with other services that can address other family needs. This conclusion is an additional reason for consolidating stronger connections between family members and community resources so that home visitors do not seek to do it all. Home

visitation programs that focus on limited, clear, well-defined, and realistic objectives have the greatest chance of success by enabling program staff to sustain program focus and to use limited resources to achieve realistic expectations (General Accounting Office, 1990). These concerns with the effective implementation of home visitation programs are familiar in light of the foregoing analysis of the challenges of social support interventions to address risks associated with social isolation. Together, they do not indicate that efforts to improve social support in the lives of troubled children and their families are not worthwhile. In fact, rigorous evaluation of promising programs suggests their potential, even for long-term benefits (e.g., Olds, 2006). The concerns do, however, indicate that it is crucial to move beyond a general expectation that providing social support in itself will yield many benefits to recipients. Instead, in recognition that (a) social support is multifaceted; (b) potential recipients have diverse needs, expectations, and personal and cultural backgrounds; (c) their natural social networks have unique resources and difficulties; (d) their communities likewise have unique constellations of material and human capital; (e) support providers have needs for training and support that are central to an effective intervention; and (f) social support cannot alone address the complex needs of recipients, the importance of clear thinking concerning the purposes of social support interventions, how these objectives should be translated into program strategies, and how the outcomes of these efforts should be evaluated is warranted.

Social Support and the Maintenance of Developmental Psychopathology Psychological problems have multifaceted origins, of course, arising from the interaction of biological, ecological, cognitive, and other influences. The absence of social support, especially in the context of conflict in close relationships, adds further risk to the development of internalizing and externalizing disorders. The availability of social support can, in turn, help to buffer the onset of clinical symptomatology. Social support and its absence are relevant also to the persistence of symptomatology over time. In other words, once clinical problems have developed, their maintenance may be associated with continuing social adversity and social isolation. Several research literatures suggest how this may be true. Children with anxiety disorders are highly vigilant for and hyperresponsve to situations associated with fearful stimuli. They interpret everyday situations in ways that exaggerate potential threat, are acutely sensitive to their own visceral signs of fear arousal, and become preoccupied with their negative emotion (Vasey & Ollendick, 2000). They are, not surprisingly, challenging for caregivers to help, but research evidence indicates that the efforts of parents to be emotionally supportive may exacerbate rather than alleviate anxious symptomatology (Thompson, 2001; Vasey & Dadds, 2001). Many parents of anxious children respond sympathetically and protectively to the fear expressed by their offspring, assisting the child in avoiding the fear-provoking event but, as a consequence, offering few opportunities to master the anxiety (Dadds, Barrett, Rapee, & Ryan, 1996; Gerlsma, Emmelkamp, & Arrindell, 1990). In light of studies showing an intergenerational family history for anxiety disorders and their high heritability, these characteristics of parent-child interaction suggest that anxiety is

learned in families as part of the shared environment and that parents may become anxious in situations in which their offspring are also fearful (Eley, 2001; Hettema, Neale, & Kendler, 2001). Thus, a child who responds to an anticipated encounter with a fear-evoking event with screaming, tantrums, hiding, and aggressive resistance offers powerful incentives for adults to accede to the child's fears. If the adults do so and the child subsequently calms down, each partner is negatively reinforced for behavior that helps to perpetuate anxious symptomatology (Vasey & Ollendick, 2000). At the same time, parents may be worried about the effects of anxious pathology on the child's capacities to act in a socially and developmentally appropriate manner, and thus parental criticism and possible rejection may accompany parental overprotectiveness (Gerlsma et al., 1990). Thus, childhood anxiety disorders are likely to be accompanied by troubled family relationships. They are troubled by aversive encounters focused on the child's efforts to avoid fear-provoking events that are inadvertently reinforced by parents' efforts to be emotionally supportive and helpful. They are troubled also by the parents' mixed response to the child's behavior—overprotective but also critical—that contributes to parent-child relationships of insecurity and uncertainty. Parental behavior in these instances is well intentioned: Adults are striving to provide social support, even though their efforts inadvertently reinforce anxious symptomatology in children (Thompson, 2001). One reason for family difficulty is that the emotional support offered by parents does not contribute to alleviating anxious fear. The problem in these circumstances is that parental social support is conceived narrowly as emotionally supportive behaviors that, paradoxically, maintain rather than reduce child anxiety. A broader conceptualization of social support is needed that incorporates parental practices that help children to progressively cope with their fearful emotion, with parental assistance, that children may initially experience as distressing but which ultimately contribute to a reduction in child anxiety. Another illustration of the influence of social support in the maintenance of clinical problems is the well-developed research literature linking peer relationship problems and childhood loneliness (see Asher, Parkhurst, Hymel, & Williams, 1990, for a review). Several studies have explored the associations among peer rejection, loneliness, and depressive symptoms in children (e.g., Boivin, Hymel, & Bukowski, 1995; Boivin, Poulin, & Vitaro, 1994; Burks, Dodge, & Price, 1995; Nangle, Erdley, Newman, Mason, & Carpenter, 2003; Oldenburg & Kerns, 1997). Nangle and his colleagues (2003), for example, found that a fully mediational model of the influence of peer relationships on children's loneliness and depression was warranted in their study of 193 third- through sixth-grade boys and girls. In this study, loneliness was the gateway through which problems in peer relationships influenced the development of depression. However, friendships buffered children from depressive symptomatology. The influence of popularity, or peer acceptance, was completely accounted for by the fact that peer acceptance temporally preceded friendship, suggesting that more popular children were likely to have both larger friendship networks and better relative quality of peer relationships. The researchers concluded that before adolescence, “the mutuality that is unique to friendships appears to be critical” (p. 552). Peer rejection is important not only to the development of depressive symptomatology but also to its maintenance over time, as

children without friendships are likely to continue to feel lonely, self-deprecating, and isolated in their peer networks.

Social Support and the Prevention and Treatment of Developmental Psychopathology Social support is important to developmental psychopathology not only because of its contribution to understanding the etiology or maintenance of pathological symptomatology but also because of its promise for pioneering avenues for prevention and therapeutic assistance. The core features of social support—counseling and guidance, emotional nurturance, information, skill acquisition, and sometimes material aid—are components of successful therapeutic efforts in all theoretical approaches. Whether in the context of individual therapy sessions, peer counseling, group therapy, parent education or parent support groups, therapeutic preschool programs, crisis counseling, or other therapeutic avenues, social support is an almost inescapable element of successful clinical intervention. Understanding the nature of social support as well as obstacles to its efficacy in promoting psychological well-being and recognizing conditions that foster perceptions of support in troubled individuals are each important to successfully enlisting social support in therapeutic efforts. Moreover, just as social support is incorporated into most forms of psychological treatment, it is also a central feature of prevention efforts to avert psychological problems or to avoid their recurrence, especially for individuals at risk. Social support is a contributor to long-term adaptive functioning as well as immediate assistance to individuals and families in need. Despite its ubiquitous contribution to therapeutic endeavors, incorporating social support in prevention and intervention strategies presents significant challenges. The forms of social support that are most helpful to individuals experiencing psychological distress are not selfevident, and, as we have seen, not all social support efforts are effective in achieving therapeutic or preventive goals. These dual observations are crucial considerations because of the many ways that well-intentioned supportive efforts can be rendered ineffective in changing maladaptive behaviors, fostering psychological well-being, or accomplishing therapeutic goals. Earlier in this chapter, we considered examples of how the emotional support of family and friends did not prevent troubled mothers from committing fatal child abuse and how parental efforts to respond sympathetically and protectively does not promote effective coping in offspring with anxiety disorders. For social support to be effective, its purposes and functions must be strategically considered within a broader array of therapeutic efforts, keeping in mind that social support should include not only emotional sustenance and counseling but also reeducation, behavioral change, and monitoring the well-being of those it is intended to assist. In the next section, we consider five interventions that incorporate aspects of social support for purposes of prevention or therapeutic remediation. These programs were chosen to sample a broad variety of intervention strategies and goals, ranging from efforts focused on strengthening informal support through peers or mentoring relationships to more formal therapeutic strategies to alter family functioning or mental representations of relational support to changes in child

welfare practices to improve the well-being of foster children. The outcomes assessed in the evaluations of these interventions are equally diverse, ranging from behavioral competence to attachment security to cortisol regulation. Together, they illustrate the opportunities and the challenges of enlisting social support for preventive or therapeutic purposes. They also illustrate how far we have to go in the evaluation of promising intervention strategies, the science underlying program design, and the extension of that science in conceptualizing the provision of social support and its anticipated benefits. Peer Support Programs As noted earlier, one strategy for enlisting social support to remediate clinical problems is to engage natural helpers. For children and adolescents, doing this usually means providing training and supervision to peers who are then expected to help their classmates or friends. The goals of such strategies may be to increase the total number of peers who interact with the target child or to teach target children social skills that will help them to attract more friends. Enlisting healthy, socially skilled peers to help children or adolescents who have behavioral or emotional disorders or who are at risk for the development of such disorders has intuitive appeal because of the potential to address both of those goals. However, as we have noted, what appears simple and straightforward is often deceptively so. Lewis and Lewis (1996) identified several concerns with involving peers in helping each other. Most notably, in programs without careful role definitions and professional supervision, young people who are motivated by a desire to help others may find themselves in situations that require significantly more training, expertise, and maturity than they have. Lewis and Lewis focused their analysis on risks to peer helpers in programs enlisting peer support for children at suicidal risk, and they reported findings from a descriptive study of Washington schools indicating that suicide rates were higher at schools where peer helpers were not supervised by professional counselors. As we have noted earlier, the potential advantages of enlisting natural supporters to aid troubled children or families must be balanced by their limitations in expertise and capability. One of the lessons of the social support literature is that designing interventions with carefully defined target behaviors and narrowly focused strategies is likely to avoid some of the ambiguity of more general supportive or mentoring approaches. Such interventions also provide more precise guidance about what works and what does not work. Consistent with this view, one group of researchers (Christopher, Hansen, & MacMillan, 1991; Guevremont, MacMillan, Shawchuck, & Hansen, 1989) trained 7- to 9-year-old children in specific social skills and rewarded them for playing with identified classmates. Same-sex peers were trained in initiating, responding to refusals, maintaining interactions, and responding constructively to negative behavior. Each was then paired with a socially isolated, dysphoric child and rewarded for playing with that child during one daily recess. The intervention resulted in a meaningful increase in positive social interactions for the target children. The improvement in peer interactions for the target children occurred with both the designated helpers and with other classmates, and the levels of positive interactions were comparable to those of social comparison children in the same setting. The treatment gains increased at a 4-month follow-up

(Guevremont et al., 1989). In a similar study that also targeted socially isolated, dysphoric children, the researchers increased the training for peer helpers and achieved similarly promising results. Gains in positive interactions were maintained at the 4-month follow-up, and the positive effects generalized to situations when the intervention was not used. In addition, there was no evidence of negative impact of social support on the helpers (Christopher et al., 1991). A similar approach has proven successful with high-risk preschool children. Fantuzzo and colleagues have demonstrated that their Resilient Peer Treatment (RPT) program improves the social competence of maltreated and nonmaltreated socially withdrawn children in Head Start programs. In RPT, highly socially competent peers help to create regular positive peer play experiences, with interactions supported by a trained family volunteer play supporter. Supporters created a space for the two children to play together and provided guidance and encouragement for interactive play. The intervention involved 15 play sessions over a 2- month period. Multiple experimental evaluations of this program indicated improvements in withdrawn children's teacher-rated social competence and decreases in withdrawn behavior within sessions (Fantuzzo et al., 1996) and with all peers during classroom free play sessions 2 weeks later (Fantuzzo, Manz, Atkins, & Meyers, 2005). Teachers also rated children as higher in self-control and lower in disconnected and disruptive play compared with control group children. Longer-term follow-up is necessary, of course, but successful intervention enlisting peers to help young high-risk children is promising, given its potential for improving later trajectories of social interactions and behavior problems. Taken together, these programs illustrate the potential benefits that may be derived from enlisting peers as informal social support agents for children who are at risk of developmental problems or face significant adversity. They also illustrate, however, the need for clarity of goals that are suitable to the enlistment of peer helpers and the importance of guidance and support of children who are recruited to help their peers. Youth Mentoring Mentors outside the family may play a significant role in providing social support to youth, and this fact has given rise to a proliferation of formal mentoring programs as preventive and therapeutic interventions for at-risk youth. Over 5,000 mentoring programs were operating in the United States in 2005 (MENTOR/National Mentoring Partnership, 2006), spanning multiple program models, such as community-based programs and school- or faith-based approaches. Some programs serve children with existing emotional or behavioral problems, while others target children with one or more risk factors for developmental difficulties (e.g., children with incarcerated parents). Mentoring programs link youth in difficulty who may not have strong natural mentors with an adult or older peer to provide support and guidance and serve as a positive role model. Rhodes and colleagues (Rhodes, 2002, 2005; Rhodes & DuBois, 2008) have proposed that when mentoring relationships develop mutuality and trust, they can potentially have a positive influence on socioemotional functioning (including developing emotion regulation), cognitive capacities (such as perspective-taking skill), and identity development.

Although formal youth mentoring programs garnered enthusiasm and funding from public and private sources with little evidence of their effectiveness for many years, recent well-designed evaluations and meta-analyses provide some indication of their potential benefits for children and the mediators of these benefits. One example is the Big Brothers/Big Sisters (BBBS) impact study (Grossman & Tierney, 1998; Rhodes, 2002). The BBBS community-based program individually matches adult mentors to children from single-parent homes. Mentors commit to at least one contact per week for a minimum of 1 year, with the expressed goal of developing a friendship with the child. The program provides high levels of supervision and ongoing support of matches. In the experimental evaluation study, diverse youth ages 10 to 16 were randomly assigned either to a treatment group and matched as quickly as possible with a mentor (n = 571) or to a waitlist control group (n = 567). Many of the children in the study had experienced relationship disturbances associated with single parenting, including parental divorce or death, domestic violence, and abuse. At the 18-month follow-up, 378 treatment group youth had been matched with a mentor for 11 months on average. At follow-up, youth who had a mentor were significantly less likely to have initiated drug or alcohol use and reported fewer instances of physically aggressive behavior. Mentored youth also reported greater trust and communication with parents and more emotional support from peers than did control group youth. The latter findings appear to support program claims that the presence of one supportive, healthy relationship may carry over into other relationships by increasing social skills, constructive forms of emotion regulation and the ability to trust others. If this is the case, an effective mentoring relationship may enhance social support both directly and indirectly, with positive implications for overall perceptions of social support and mental health (see also Rhodes, Reddy, & Grossman, 2005). Recent meta-analyses conclude that effects of youth mentoring programs are relatively consistent but quite modest. In a 2011 meta-analysis covering 73 evaluation studies published after 1998, DuBois and colleagues reported effect sizes of .17 for positive social/relational outcomes, .15 for positive psychological/emotional outcomes, and .21 for lower conduct problems (DuBois, Portillo, Rhodes, Silverthorn, & Valentine, 2011). These results were consistent with those of an earlier analysis (DuBois, Holloway, Valentine, & Cooper, 2002). Although very few studies reported on longer-term effects, when follow-up data were available (7 studies), the effects of program participation showed little to no deterioration over periods lasting up to 4 years. DuBois and colleagues also reported a number of moderators of mentoring program effectiveness. Prominent among these were characteristics of the mentor-child relationship. Several characteristics were important to program efficacy. First, the positive effects of mentoring relationships were more likely to exist in longer, more established relationships that provided extended opportunities for the development of trust. For example, in a reanalysis of the BBBS evaluation data, Rhodes and colleagues (2005) found that the positive effects of mentors on parent-child relationships and on lowering levels of substance use were present only in relationships established for at least 1 year. Second, the style of mentoring also plays a role in the efficacy of this helping relationship. Reanalyzing the BBBS data, Langhout, Rhodes, and Osborne (2004) found that there were different mentoring relationship styles characterized

by varying levels of support, structure, and activity, and these were differentially related to youth outcomes at follow-up. An “unconditionally supportive” mentoring style, for example, was high in levels of positive support but lower in structure of conversations concerning goals and the number of shared activities. By contrast, a “moderate” mentoring style was comparable in shared activities and structured conversations but lower in positive support, while an “active” style was characterized by the highest number of shared activities. Interestingly, youth experiencing “moderate” mentoring showed the best outcomes on follow-up, with decreased feelings of alienation from parents, lower levels of peer conflict, and improved academic competence compared to controls. By contrast, an “unconditionally supportive” mentoring style was associated with an increase in both alienation and peer conflict over the study period. The authors suggested that children in this program may have benefited most from mentors who provided a sense of structure and goal orientation within the context of an emotionally positive adult-child relationship rather than unconditional positive regard with little structure. Are there groups of children who are more likely to benefit from the social support provided by a mentor? Although effects held across a range of ages, from early childhood through late adolescence (DuBois et al., 2011), other child characteristics moderated effects. For example, children with prior behavior problems and children who brought either high individual or environmental risk appeared to benefit more from mentoring. It is important to note that mentoring programs appear most effective for those children who face an intermediate but not an overwhelming degree of challenges; as earlier noted, the potential impact of nonprofessional mentors may be blunted by the complex problems of children facing both high individual and environmental challenges. The nature of individual mentoring programs may also bear on their effects. DuBois et al. (2002, 2011) reported that mentoring effects were considerably stronger when programs were structured using research-based best practices, such as matching mentors and children based on individual interests, family involvement, and high levels of program supervision and mentor support. Taken together, studies of the efficacy of formal and informal mentoring relationships for enhancing the well-being of at-risk children and youth suggest that these programs can be helpful for improving social functioning, reducing behavioral problems, and even strengthening academic success. They suggest that in many circumstances, the positive relationship established with an older mentor may provide a foundation for improved relationships with family members, peers, and others in the child's social ecology and may contribute to enhancing perceptions of social support that are important to psychological well-being. The findings also indicate, however, that understanding the characteristics of effective mentoring relationships is important, especially as they concern the match between youth and mentors in background, interests, and other characteristics; the nature and longevity of their relationship; and the severity of youth difficulties that a mentoring relationship is expected to address. Family Support Intervention Interventions designed particularly for family members, especially parents, are often part of the therapeutic effort for children and adolescents. Families are the primary source of support for

children, and family dysfunction—whether it is manifested in child maltreatment, domestic violence, serious marital conflict, parental depression, or other conditions—poses significant risk to children's emotional health. Consequently, the quality and effectiveness of the support that family members provide to children with clinical problems is crucial to children's healthy adjustment. As we have seen, however, not all efforts to enhance social support are helpful, and, indeed, emotional support alone may not be effective in remediating children's problems in many situations. In fact, a parent's emotional engagement may have negative effects in certain circumstances. For example, children's emotional or behavioral difficulties may lead to heightened criticism and hostility toward the child (Hooley & Richters, 1995). Expressed emotion (EE) (Hirshfield, Biederman, Brody, Faraone, & Rosenbaum, 1997a) is an index of parental attitudes of criticism or emotional overinvolvement in a child's problems that has been studied as a contributor to the onset, maintenance, or relapse of a number of clinical problems, including schizophrenia (Brown, Birley, & Wing, 1972), depression (Hooley, Orley, & Teasdale, 1986), bipolar disorder (Pavuluri, 2008), behavioral inhibition (Hirshfield et al., 1997a, 1997b), and conduct disorder (Caspi et al., 2004; Rogosch, Cicchetti, & Toth, 2004). EE is associated with these clinical outcomes because it constitutes a major relational stressor in family experience that undermines psychological functioning in an environment that should ordinarily provide social support. Family support or treatment components have been designed to include a specific focus on EE as a way of improving the quality of social support for young people with clinical problems. The RAINBOW treatment protocol, for example, is a child- and family-focused cognitive behavioral therapy for children with bipolar disorder that addresses the intense personal demands of raising a child with bipolar disorder in an effort to decrease the potentially harmful effects of EE (Pavuluri, 2008; Pavuluri et al., 2004). Elements of this program include encouraging family members to distinguish helpful from unhelpful reactions in their efforts to cope with a child who can be difficult to live with, helping parents model appropriate strategies for affect regulation, fostering shared effective problem-solving strategies in which parents and target children jointly participate, assisting children in their efforts to develop successful peer relationships, and identifying other sources of social support. The program, designed for families with children in the middle childhood or adolescent years, consists of twelve 60-minute sessions that occur weekly over a 3-month period. Conclusions from a study of 34 families participating in this program indicated that symptom severity for children decreased significantly following therapy and parents reported strong satisfaction with the treatment, although there were no family-based measures of the emotional environment (Pavuluri et al., 2004; see also West et al., 2009). A maintenance model of this intervention showed that these benefits persisted 3 years after initial participation in the program (West, Henry, & Pavuluri, 2007). Family support interventions like this illustrate how informal social support in the family can be strengthened with the therapeutic assistance of formal helpers from the psychological and psychiatric community. This collaboration can occur even in circumstances when the child's symptom severity has contributed to elevated family stress and hostility toward the child. The

duration of the program, however, and the need for a maintenance regimen together illustrate that these benefits are not easily achieved in troubled families. Child-Parent Psychotherapy Because social support is based on the quality of helping relationships, efforts like family support interventions are designed to strengthen these relationships by altering the behavior of each partner. But often behavioral change is insufficient because relational interactions are affected not only by the behavior of the immediate partner but also by the individual's prior history of the relationship and its psychological impact. This idea is captured in attachment theory in the concept of internal working models, or mental representations that emerge from early caregiving experience and subsequently color the child's expectations and interactions for subsequent relational partners (Thompson, 2006, 2008). For adults, the quality of their earlier close relationships can color their capacity to respond to a child's needs and emotions. The importance of these mental representations of relationships is reflected also in the social support literature reviewed earlier, particularly in the realization that perceived support is more important to psychological well-being than are the actual experiences of receiving support from social network associates. The expectation that others are available to provide assistance, which is important to a sense of social support, may be undermined by earlier experiences of trauma, neglect or abuse, relational violence, and other experiences that impair how individuals are capable of functioning in other relationships. One approach to therapeutic assistance is thus to address the relational representations that may be obstacles to functioning effectively in potentially supportive relationships. One example is Child-Parent Psychotherapy, an attachment-based approach designed for young children who have experienced a traumatic event, such as maltreatment or exposure to domestic violence, that can be manifested in posttraumatic symptomatology and attachment disorganization (Lieberman, Ghosh Ippen, & Van Horn, 2006; Lieberman & Van Horn, 2008; Lieberman, Van Horn, & Ghosh Ippen, 2005). This is a two-generation intervention because of the recognition that parents of children in these circumstances often have their own history of traumatic experiences that significantly affect their interactions with the child. Thus, traumafocused therapy for the parent is also a focus of intervention, which may include recollections of difficult childhood experiences as well as explicit attention to the impact of these experiences on the mother's representations of herself and her child. In addition, therapy focuses on parent-child interaction, including guidance for the mother concerning the child's developmental needs and fostering healthier patterns of maternal behavior to alter the young child's expectations for maternal responsiveness and encourage the development of a more secure attachment to her. Age-appropriate therapeutic assistance to the child is also a component of the intervention. Evaluation studies of Child-Parent Psychotherapy have found a reduction in behavior problems and posttraumatic stress symptomatology in preschoolers in a therapeutic group compared to a control group (Lieberman et al., 2006). In an adaptation of this approach, Toth and her colleagues (2002) used Preschooler-Parent Psychotherapy with a sample of young children who had been maltreated by their mothers and focused on changes in children's representations of the mother and their relationship (assessed

through semiprojective narrative story stem completions with dolls) in the assessment of therapeutic outcome. Children and their mothers were seen in weekly 60-minute sessions with a clinician over the course of 12 months. Toth and colleagues found that children in the intervention group exhibited a greater decline in negative maternal representations, a greater decrease in negative self-representations, and more positive expectations of the mother-child relationship by postintervention compared to children treated in alternative approaches. In another study, Toddler-Parent Psychotherapy was used with a sample of mothers with major depressive disorder and their children. After a year of weekly therapeutic sessions, the rate of secure attachment for children in the intervention group had increased significantly and was higher than for a nontreated control group and a nondepressed community group (Toth, Rogosch, Manly, & Cicchetti, 2006; see also Cicchetti, Toth, & Rogosch, 1999). Similar findings have been reported in other applications of this psychotherapeutic approach. Cicchetti, Rogosch, and Toth (2006), for example, reported significant increases in secure attachment in 2-year-olds after participating with their mothers in a year of Infant-Parent Psychotherapy. In a follow-up study, Stronach, Toth, Rogosc, and Cicchetti (2013) reported that these benefits were sustained for a year following the intervention. Taken together, these studies underscore the importance of attention to the representational dimensions of receiving support in the design of social support interventions. Indeed, they suggest that without consideration of how prior experience has affected a potential recipient's expectations for assistance from others, efforts to enhance social connections or aid may be blunted because greater received support may not yield increased perceptions of support or expectations that others will be helpful in the future. Conversely, support interventions that focus explicitly on improving expectations of support from others, especially in the context of efforts to strengthen supportive social network ties, may provide greater benefits. But much more remains to be understood about how mental representations of relationships change and grow over time and, in particular, how enduring are therapeutic interventions that address these representations related to social support. Supportive Interventions in Foster Care: Behavioral and Biological Outcomes A theme of the research surveyed in this chapter is that stress has psychological and physiological consequences, and the social buffering of stress is likewise psychological and biological in nature. Social support is believed to foster more effective coping and promote psychological well-being through its effects on an individual's psychological stress and its physiological substrates. Although research confirms this multilevel orientation, supportive programs rarely examine the effects of interventions on biological processes, such as stress neurobiology, that are believed to be associated with stress and coping. Even more rarely do they consider concurrent behavioral and biological outcomes of social support. A final illustration of a social support intervention that does so shows how these biological and behavioral outome can be studied together and provides a model for similar kinds of evaluation research in the future. Children in foster care commonly experience chronic adversity owing to the conditions associated with their initial removal from the family, their experience in foster care, and

transitions between care settings. Considerable research has shown that these experiences create stress that is manifested in behavioral problems, distrustful relationships, and dysregulated patterns of biological stress reactivity. Concerning the latter, children in foster care have been shown to exhibit either elevated diurnal cortisol levels or flattened diurnal cortisol, depending, it appears, on the quality of care they have received (e.g., abusive or neglectful treatment) (Bruce, Gunnar, Pears, & Fisher, 2013). In either case, their biological stress systems, most notably HPA activity, have become dysregulated as the result of their experiences of care. This dysregulation of biological stress reactivity is consistent with findings for other groups of children experiencing chronic stress, such as those who are maltreated or are in institutional care (see Thompson, 2014a, 2015, for reviews). Can social support provide benefits to children in foster care that improve their behavioral functioning and normalize biological stress reactivity? The results of two intervention studies that share many common features suggest that this may be possible. In one program, Fisher and colleagues (Fisher, Stoolmiller, Gunnar, & Burraston, 2007; Fisher, Van Ryzin, & Gunnar, 2011) designed an intervention to ease young children's transition to new foster care placements and enhance continuity of care to reduce the stresses of such placement. Before the intervention began, these children showed a profile of flattened cortisol activity as a result of their earlier placements. The intervention program was designed to promote warm, responsive, and consistent relationships between children and new foster parents in which positive behavior was encouraged, problem behavior was reduced, and caregiver stress was lowered. The program included individualized sessions with child therapists, weekly playgroup sessions, and other child-focused services. Foster parents completed intensive training prior to the children's placement, and they continued to receive support and supervision in daily phone contacts, weekly group meetings, and on-call assistance. The program was thus a 2-generation intervention involving multiple adults who functioned as caregivers for the child. Compared to children assigned to regular foster care placements, who showed the pattern of cortisol hyporesponsiveness described earlier, children in the intervention group showed patterns of HPA reactivity that progressively resembled the normal patterns of a nonabused community comparison group over the course of 6 to 12 months of treatment. Reductions in foster parent stress were directly implicated in these changes in young children's HPA function recovery (Fisher & Stoolmiller, 2008). In another intervention with very young children, Dozier and her colleagues (Dozier, Peloso, Lewis, Laurenceau, & Levine, 2008; Dozier et al., 2006) reported that following a home-based 10-week program, infants and toddlers in foster care showed more typical daily patterns of HPA activity and moderated cortisol reactivity to a stressor compared to a group of foster care infants in a different treatment program. This attachment-based intervention was designed to improve young children's relationships and self-regulation by assisting foster parents in better interpreting and responding to infants' signals, enhancing affectionate behavior, and providing more reliable support for infants' self-regulation. In both intervention studies, moreover, the researchers also obtained concurrent measures of behavioral change that may be viewed as additional indicators of treatment efficacy. In the Dozier study, infants and toddlers showed greater evidence of attachment to their foster parents

(Dozier et al., 2009); in the Fisher intervention, a similar gain in secure attachment behavior was noted, together with greater placement success (Fisher, Burraston, & Pears, 2005; Fisher & Kim, 2007). In light of the goal of each intervention to improve the quality of children's relational experience, and in view of research indicating disturbance of biological stress systems in young children with insecure attachments (Spangler & Grossmann, 1993), these behavioral changes should be regarded as indicators of intervention efficacy, parallel to the changes in HPA reactivity occurring over the same period. Both of these intervention programs for foster children can be regarded as focused on improving social support. Each emphasized strengthening young children's relational security by improving caregiver responsiveness and reducing caregiver stress. The success of each program in improving the quality of children's attachments while also regularizing their HPA functioning is consistent with the research reviewed earlier concerning the social buffering of physiological stress reactivity and provides some of the earliest evidence for how social support interventions can have both psychological and biological consequences. Even so, much more remains to be understood. Longer-term follow-up research on children in intervention and comparison groups is needed, for example, to examine how enduring are the changes in stress reactivity resulting from these interventions or, more specifically, the conditions in which these changes endure or not. Recognizing that the plasticity of HPA functioning can be limited in conditions of highly aversive care (e.g., Gunnar, Morison, Chisholm, & Schuder, 2001), it will also be important to understand how early adversity is related to the intensity and duration of intervention efforts required to normalize children's stress reactivity and its behavioral correlates. In the end, what makes this research field so exciting is the opportunity to extend research on socially supportive interventions to the biological domain and, in so doing, to better comprehend the behavior-biology associations that underlie psychological functioning and its changes over time. Taken together, the preventive and therapeutic programs surveyed in this section certainly do not exhaust the range of interventions that have been designed to strengthen social support in the lives of troubled individuals and families. But they illustrate the significant opportunities that exist to improve informal and formal sources of support for beneficial purposes as well as the challenges of enlisting social support for these purposes that have been discussed throughout this chapter. More specifically, these programs illustrate and confirm the interim conclusions about social support identified earlier from the research literatures discussed in the opening sections of this chapter. They show, for example, the importance of recognizing how relationships influence other relationships and the value of multigenerational interventions that provide support both to children and their caregivers. They illustrate the significance of coordinating informal and formal sources of social support for providing multifaceted avenues of assistance to those who need it. These programs show both that parents are central to children's environment of relationships but that they are not alone influential, and sometimes support can be provided by enlisting other people in children's natural social networks (e.g., peers) or by enlisting others outside the family (e.g., mentors). These programs indicate how representations of support from others are important to psychological well-being and that sometimes interventions with a representational focus can be most valuable. Most important,

these programs together show that social support efforts will be more effective when there is clarity concerning the goals they are intended to achieve. Taken together, these interventions illustrate both the lessons learned and to be applied in future social support interventions.

Conclusions and Future Directions Two of the central conclusions from the research on social support and developmental psychopathology are how complicated is the provision of social support in intervention and how beneficial everyday social support is for psychological functioning. When people are surrounded by natural networks of family, friends, neighbors, coworkers, and others who offer emotional guidance, instrumental aid, and monitor well-being, the odds in favor of psychological health and healing are meaningfully improved. But it is difficult to create the benefits of natural forms of social support in preventive or therapeutic interventions. This conclusion owes to the challenges of instituting formal helping relationships that can offer social support and also because when natural social networks are not functioning supportively (e.g., because they are drained or stressed), it is difficult to reconstitute them in healthy and helpful ways. Added to these challenges are the characteristics of support recipients and how their personal and ecological conditions can pose obstacles to the success of social support interventions. Moreover, in order to be effective in addressing clinical psychopathology, social support must accomplish more than simply providing the emotional reassurance that recipients might want most; it must also change behavior and thinking. In short, the benefits of social support are easy to envision but difficult to implement. This finding poses a challenge to researchers, theorists, clinicians, and practitioners who, after the initial wave of enthusiasm for the preventive and therapeutic benefits of social support interventions two decades ago, must now confront the host of practical challenges to effectively implement social support in the lives of at-risk children and their families. As the summary of preventive and therapeutic program models illustrates, however, there is cause for hope. From the varieties of programs that have documented measurable and predictable benefits for recipients—and from those that have failed to do so—there is now a wealth of good ideas for crafting more carefully conceived, thoughtfully designed social support interventions that have greater promise of success. There are also some excellent program models to draw on. The research literatures we have discussed related to social support are broad, detailed, and expanding and will continue to provide new ideas about support and its effects in the years to come. Although basic conclusions concerning the benefits of social support seem to be settled (some perhaps prematurely) and the era of unbridled enthusiasm concerning the promise of social support interventions is over, we believe that there are three questions meriting further research that have the greatest promise of significantly expanding understanding of social support and its relevance to developmental psychopathology. First, why is social support a stress buffer? A vast research literature documents the association between perceived (and sometimes received) social support and measures of

coping and psychological well-being. But reviewers of this research, including us, find process explanation lacking (see also Cohen & Janicki-Deverts, 2009; Thoits, 2011). Psychological researchers typically focus on intervening processes, such as enhanced selfesteem, perceptions of self-efficacy, modeling, sense of belonging and companionship, while sociological and community researchers more commonly focus on processes like resource acquisition, networking and mentoring, and skill development. But there is no widely accepted model for why social support should enhance health and coping. As we have emphasized, the multifaceted nature of social support probably means that different constellations of intervening processes are influential with different subsets of recipients. Research on stress neurobiology and the social buffering of stress reactivity may, however, offer new perspectives to this question. In the context of behavior-biology transactions—or, as emphasized by Thompson (2015b), behavior-representation-biology interactions— researchers in developmental neuroscience are beginning to identify some of the constituent mechanisms by which social support modulates neurobiological stress reactivity. The intervening processes currently hypothesized include, for example, the effects of social support on the biological regulators of HPA activity (including prefrontal influences but possibly also limbic regulators), the downregulation of the proinflammatory tendencies induced by chronic stress, and the influence of other neurohormonal processes, including oxytocin (see Hostinar et al., 2014; Kiecolt-Glaser et al., 2010; Ulrich-Lai & Herman, 2009). This research is still at an early stage, and establishing connections between neuroscience and behavioral functioning in this area poses an additional challenge. But there is the potential for the development of theoretical models of social support that describe the ongoing transactions between different levels of the neuroaxis and behavior that can help clarify the physiological mediators between social support and mental health as well as other neurobiological processes associated with clinical psychopathology that may alter these processes. This is an exciting prospect for theoretical portrayals of biology-behavioral interactions and has potentially important intervention applications. Second, what developmental processes guide children's changing representations of social support in their environment of relationships? One of the conclusions of this chapter is that the experience of receiving social support is similar for children and adults in some ways but very different in others. As we have noted, for example, children's access to social support is significantly mediated by parents and their social networks. But this review has underscored how little is known about developmental changes in children's experience or representation of social support. This is an important question for future understanding of the association of social support and developmental psychopathology. In this discussion, for example, we have underscored that social support is developmentally constitutive for children in a way that it is not for adults: Children depend significantly on the various kinds of support they receive from others for their survival, health, and well-being. This dependency has an effect on their reactions to receiving assistance: Children do not seem vulnerable, for example, to the feelings of indebtedness, humiliation, or vulnerability that adults often experience when receiving unreciprocated assistance from another (for that matter, they are also much less likely to experience gratitude) because it is widely acknowledged that

children depend on the solicitude of adults. But we know little about how these perceptions change with increasing age and children's growing autonomy. We know little about developmental changes in how children represent more broadly the support that is available to them from parents and extended family members, neighbors, peers, teachers, and others in their social ecology. How are individual differences in characteristics such as temperament influential in the growth of these mental representations? And although the attachment literature suggests that the internal working models of secure attachment cause children to perceive greater supportiveness from others and perhaps also to elicit that support, little is known of how this occurs or whether it creates a snowball effect that might contribute to some of the broader socioemotional outcomes associated with attachment security. Finally, the emergence of online forms of social support raises important new questions concerning the extent to which children and youth experience support in these transactions and the conditions in which this support can occur. In short, better understanding of developmental processes associated with children's experience and representation of social support is a crucial future research task. Third, what are the essential ingredients for effective preventive and therapeutic social support interventions? As researchers have increasingly recognized the complexity of designing social support interventions for at-risk children and their families, a number of promising interventions have been developed, some of which we have discussed in this chapter. Furthermore, these efforts have provided clinicians with a short list of things to do (e.g., engage recipients; clarify goals; support help providers; respect culture) and to avoid (e.g., expect more than informal helpers can do; view social support alone as the answer to recipients' needs) in their thinking about intervention strategy. It is probably too soon to expect that a recipe for effective social support interventions will emerge, especially in light of how limited current evaluation studies are in clarifying these ingredients and the need for long-term follow-up studies. But such a recipe should be a long-range goal receiving current attention by those who are involved in studying, designing, conducting, and evaluating these interventions. Social support is essential to psychological growth, well-being, and effective coping. It is also a core element to successful preventive and therapeutic efforts to assist at-risk children and their families. Understanding how to better engage social support into the lives of those who need it remains an essential and enlivening research and clinical endeavor.

References Administration for Children and Families. (2002). Making a difference in the lives of children and families: The impacts of Early Head Start programs on infants and toddlers and their families. Washington, DC: U.S. Department of Health and Human Services. Anan, R., & Barnett, D. (1999). Perceived social support mediates between prior attachment and subsequent adjustment: A study of urban African American children. Developmental Psychology, 35, 1210–1222. doi: 10.1037/0012–1649.35.5.1210 Asher, S. R. Parkhurst, J. R., Hymel, S., & Williams, G. A. (1990). Peer rejection and loneliness in childhood. In S. R. Asher & J. D. Coie (Eds.), Peer rejection in childhood (pp.

253–273). New York: Cambridge University Press. Avellar, S., Paulsell, D., Sama- Miller, E., & Del Grosso, P. (2013). Home visiting evidence of effectiveness review: Executive summary. Washington, DC: Office of Planning, Research and Evaluation, Administration for Children and Families, U.S. Department of Health and Human Services. Badr, H., Acitelli, L. K., Duck, S., & Carl, W. J. (2001). Weaving social support and relationships together. In B. R. Sarason & S. Duck (Eds.), Personal relationships: Implications for clinical and community psychology (pp. 1–14). Chichester, UK: Wiley. Bagwell, C. L., Newcomb, A. F., & Bukowski, W. M. (1998). Preadolescent friendship and peer rejection as predictors of adult adjustment. Child Development, 69, 140–153. doi: 10.2307/1132076 Baker, A. J. L., Piotrkowski, C. S., & Brooks-Gunn, J. (1999). The Home Instruction Program for Preschool Youngsters (HIPPY). Future of Children, 9, 116–133. Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2006). Gene–environment interaction of the dopamine D4 receptor (DRD4) and observed maternal insensitivity predicting externalizing behavior in preschoolers. Developmental Psychobiology, 48, 406–409. doi: 10.1002/dev.20152 Bakermans-Kranenburg, M. J., van IJzendoorn, M. H., Pijlman, F. T. A., Mesman, J., & Juffer, F. (2008). Experimental evidence for differential susceptibility: Dopamine D4 receptor polymorphism (DRD4 VNTR) moderates intervention effects on toddlers' externalizing behavior in a randomized controlled trial. Developmental Psychology, 44, 293–300. doi: 10.1037/0012–1649.44.1.293 Barrera, M. (1981). Social support in the adjustment of pregnant adolescents: Assessment issues. In B. Gottlieb (Ed.), Social networks and social support (pp. 69–96). Beverly Hills, CA: Sage. Barrera, M. (1986). Distinctions between social support concepts, measures, and models. American Journal of Community Psychology, 14, 413–445. doi: 10.1007/BF00922627. Beeman, S. K. (2001). Critical issues in research on social networks and social supports of children exposed to domestic violence. In S. A. Graham & J. L. Edelson (Eds.), Domestic violence in the lives of children (pp. 219–234). Washington, DC: American Psychological Association. Belle, D. (1982). Social ties and social support. In D. Belle (Ed.), Lives in stress (pp. 133– 144). Beverly Hills, CA: Sage. Belle, D. (Ed.) (1989). Children's social networks and social supports. New York, NY: Wiley.

Belle, D., & Benenson, J. (2013). Children's social networks and well-being. In A. Ben-Arieh, F. Casas, I. Frones, F. Cases, & J. E. Korbin (Eds.), Handbook of child well-being: Theories, methods, and policies in global perspective (Vol. 4, pp. 1335–1363). Dordrecht, the Netherlands: Springer. Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135, 885–908. doi: 10.1037/a0017376 Berndt, T. J., Hawkins, J. A., & Jiao, Z. (1999). Influence of friends and friendships on adjustment to junior high school. Merrill-Palmer Quarterly, 45, 13–41. Bessière, K., Kiesler, S., Kraut, R., & Boneva, B. S. (2008). Effects of Internet use and social resources on changes in depression. Information, Communication, and Society, 11, 47–70. doi: 10.1080/13691180701858851 Blair, C., & Raver, C. C. (2012). Child development in the context of adversity: Experiential canalization of brain and behavior. American Psychologist, 67, 309–318. doi: 10.1037/a0027493 Blair, C., Raver, C. C., Granger, D., Mills-Koonce, R., Hibel, L., & the Family Life Project Key Investigators (2011). Allostasis and allostatic load in the context of poverty in early childhood. Development and Psychopathology, 23, 845–857. doi: 10.1017/S0954579411000344 Bogat, G. A., Caldwell, R. A., Rogosch, F. A., & Kriegler, J. A. (1985). Differentiating specialists and generalists within college students' social support networks. Journal of Youth and Adolescence, 14, 23–35. doi: 10.1007/BF02088644 Boivin, M., Hymel, S., & Bukowski, W. M. (1995). The roles of social withdrawal, peer rejection, and victimization by peers in predicting loneliness and depression. Development and Psychopathology, 7, 765–785. doi: 10.1017/S0954579400006830 Boivin, M., Poulin, F., & Vitaro, F. (1994). Depressed mood and peer rejection in childhood. Development and Psychopathology, 6, 483–498. doi: 10.1017/S0954579400006064 Bost, K., Vaughn, B., Washington, W., Cielinski, K., & Bradbard, M. (1998). Social competence, social support, and attachment: Demarcation of construct domains, measurement, and paths of influence for preschool children attending Head Start. Child Development, 69, 192–218. doi: 10.2307/1132080 Brown, G. W., Birley, J. L. T., & Wing, J. K. (1972). Influence of family life on the course of schizophrenic disorders: A replication. British Journal of Psychiatry, 121, 241–258. doi: 10.1192/bjp.121.3.241 Bruce, J., Gunnar, M. R., Pears, K. C., & Fisher, P. A. (2013). Early adverse care, stress neurobiology, and prevention science: Lessons learned. Prevention Science, 14, 247–256. doi: 10.1007/s11121–012–0354–6

Bukowski, W. M., Brendgen, M., & Vitaro, F. (2007). Peers and socialization: Effects on externalizing and internalizing problems. In J. E. Grusec & P. D. Hastings (Eds.), Handbook of socialization theory and research (pp. 355–381). New York, NY: Guilford Press. Caplan, S. E., & Turner, J. S. (2007). Bringing theory to research on computer-mediated comforting communication. Computers in Human Behavior, 23, 985–998. doi: 10.1016/j.chb.2005.08.003 Caspi, A., Moffitt, T. E., Morgan, J., Rutter, M., Taylor, A., Arseneault, L.,… Polo-Tomas, M. (2004). Maternal expressed emotion predicts children's antisocial behavior problems: Using monozygotic-twin differences to identify environmental effects on behavioral development. Developmental Psychology, 40, 149–161. doi: 10.1037/0012–1649.40.2.149 Cassidy, J. (2008). The nature of the child's ties. In J. Cassidy & P. R. Shaver (Eds.), Handbook of attachment (2nd ed., pp. 3–22). New York, NY: Guilford Press Cauce, A. M., Felner, R. D., & Primavera, J. (1982). Social support in high-risk adolescents: Structural components and adaptive impact. American Journal of Community Psychology, 10, 417–428. doi: 10.1007/BF00893980 Cauce, A. M., Reid, M., Landesman, S., & Gonzales, N. (1990). Social support in young children: Measurement, structure, and behavioral impact. In B. R. Sarason, I. G. Sarason, & G. R. Pierce (Eds.), Social support: An interactional view (pp. 64–94). New York, NY: Wiley. Ceballo, R., & McLoyd, V. (2002). Social support and parenting in poor, dangerous neighborhoods. Child Development, 73, 1310–1321. doi: 10.1111/1467–8624.00473 Christopher, J. S., Hansen, D. J., & MacMillan, V. M. (1991). Effectiveness of a peer-helper intervention to increase children's social interactions: Generalization, maintenance, and social validity. Behavior Modification, 15, 22–50. doi: 10.1177/01454455910151002 Cicchetti, D., Rogosch, F. A., & Toth, S. L. (2006). Fostering secure attachment in infants in maltreating families through preventive interventions. Development and Psychopathology, 18, 623–649. doi: 10.10170S0954579406060329 Cicchetti, D., & Toth, S. (2006). Developmental psychopathology and preventive intervention. In W. Damon & R. M. Lerner (Eds.), Handbook of child psychology, 6th ed., Vol. IV: Child psychology in practice (K. A. Renninger & I. E. Sigel, Vol. Eds.) (pp. 497–547). Hoboken, NJ: Wiley. Cicchetti, D., Toth, S. L., & Rogosch, F. A. (1999). The efficacy of toddler-parent psychotherapy to increase attachment security in offspring of depressed mothers. Attachment & Human Development, 1, 34–66. doi: 10.1080/14616739900134021. Cicchetti, D., & Valentino, K. (2006). An ecological-transactional perspective on child maltreatment: Failure of the average expectable environment and its influence on child development. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, 2nd ed.,

Vol. III: Risk, disorder, and adaptation (pp. 129–201). Hoboken, NJ: Wiley. Coan, J. A., Schaefer, H. S., & Davidson, R. J. (2006). Lending a hand: Social regulation of the neural response to threat. Psychological Science, 17, 1032–1039. doi: 10.1111/j.1467– 9280.2006.01832.x Coatsworth, J. D, Pantin, H., McBride, C., Briones, F., Kurtines, W., & Szapocznik, J. (2002). Ecodevelopmental correlates of behavior problems in young Hispanic females. Applied Developmental Science, 6, 126–143. doi: 10.1207/S1532480XADS0603_3 Cochran, M. (1990). Personal networks in the ecology of human development. In M. Cochran, M. Larner, D. Riley, L. Gunnarsson, & C. R. Henderson (Eds.), Extending families: The social networks of parents and their children (pp. 3–32). New York, NY: Cambridge University Press. Cochran, M. M., & Brassard, J. A. (1979). Child development and personal social networks. Child Development, 50, 601–616. doi: 10.2307/1128926 Cochran, M., Gunnarsson, L., Grave, S., & Lewis, J. (1990). The social networks of coupled mothers in four cultures. In M. Cochran, M. Larner, D. Riley, L. Gunnarsson, & C. R. Henderson (Eds.), Extending familiar: The social networks of parents and their children (pp. 86–104). New York, NY: Cambridge University Press. Cochran, M., Larner, M., Riley, D., Gunnarsson, L., & Henderson, C. R. (Eds.) (1990). Extending families: The social networks of parents and their children. New York, NY: Cambridge University Press. Cochran, M., & Niegro, S. (1995). Parenting and social networks. In M. H. Bornstein (Ed.), Handbook of parenting. Vol. 3: Status and social conditions of parenting (pp. 393–418). Hillsdale, NJ: Erlbaum. Cohen, S. (2004). Social relationships and health. American Psychologist, 59, 676–684. doi: 10.1037/0003–066X.59.8.676 Cohen, S., & Janicki-Deverts, D. (2009). Can we improve our physical health by altering our social networks? Perspectives on Psychological Science, 4, 375–378. doi: 10.1111/j.1745– 6924.2009.01141.x Cohen, S., & Wills, T. A. (1985). Stress, social support, and the buffering hypothesis. Psychological Bulletin, 98, 310–357. doi: 10.1037/0033–2909.98.2.310 Coleman, J. C. (2011). The nature of adolescence (4th ed.). London, UK: Routledge. Collins, A. H., & Pancoast, D. L. (1976). Natural helping networks: A strategy for prevention. Washington, DC: National Association of Social Workers. Cotterell, J. L. (1986). Work and community influences on the quality of child rearing. Child Development, 57, 362–374. doi: 10.2307/1130592.

Coyne, J. C., Burchill, S. A. L., & Stiles, W. B. (1991). An interactional perspective on depression. In C. R. Snyder & D. O. Forsyth (Eds.), Handbook of social and clinical psychology: The health perspective (pp. 327–349). New York: Allyn & Bacon. Coyne, J. C., & Downey, G. (1991). Social factors and psychopathology: Stress, social support, and coping processes. Annual Review of Psychology, 42, 401–425. doi: 10.1146/annurev.ps.42.020191.002153 Criss, M. M., Pettit, G. S., Bates, J. E., Dodge, K. A., & Lapp, A. L. (2002). Family adversity, positive peer relationships, and children's externalizing behavior: A longitudinal perspective on risk and resilience. Child Development, 73, 1220–1237. doi: 10.1111/1467–8624.00468 Dadds, M. R., Barrett, P. M., Rapee, R. M., & Ryan, S. (1996). Family process and child psychopathology: An observational analysis. Journal of Abnormal Child Psychology, 24, 715–734. doi: 10.1007/BF01664736 Danese, A., & McEwen, B. (2012). Childhood experiences, allostasis, allostatic load, and age-related disease. Physiology & Behavior, 106, 29–39. doi: 10.1016/j.physbeh.2011.08.019 Daro, D. (1988). Confronting child abuse. New York, NY: Free Press. Daro, D. A., & Harding, K. A. (1999). Healthy Families America: Using research to enhance practice. Future of Children, 9, 152–176. Davies, P. T., Harold, G. T., Goeke-Morey, M. C., & Cummings, E. M. (2002). Child emotional security and interparental conflict. Monographs of the Society for Research in Child Development, 67 (serial no. 270). doi: 10.1111/1540–5834.00205 Davies, P. T., & Woitach, M. J. (2008). Children's emotional security in the interparental relationship. Current Directions in Psychological Science, 17, 269–274. doi: 10.1111/j.1467–8721.2008.00588.x Dilworth-Anderson, P., & Marshal, S. (1996). Social support in its cultural context. In G. R. Pierce, B. R. Sarason, & I. G. Sarason (Eds.), Handbook of social support and the family (pp. 67–79). New York, NY: Plenum Press. Dishion, T. J., & McMahon, R. J. (1998). Parental monitoring and the prevention of child and adolescent problem behavior: A conceptual and empirical formulation. Clinical Child and Family Psychology Review, 1, 61–75. doi: 10.1023/A:1021800432380 Dozier, M., Lindheim, O., Lewis, E., Bick, J., Bernard, K., & Peloso, E. (2009). Effects of a foster parent training program on young children's attachment behaviors: Preliminary evidence for a randomized clinical trial. Child and Adolescent Social Work Journal, 26, 321–332. doi: 10.1007/s10560–009–0165–1 Dozier, M., Peloso, E., Lewis, E., Laurenceau, J.-P., & Levine, S. (2008). Effects of an attachment-based intervention on the cortisol production of infants and toddlers in foster care.

Development and Psychopathology, 20, 845–859. doi: 10.1017/S0954579408000400 Dozier, M., Peloso, E., Lindheim, O., Gordon, M. K., Manni, M., Sepulveda, S., & Ackerman, J. (2006). Developing evidence-based interventions for foster children: An example of a randomized clinical trial with infants and toddlers. Journal of Social Issues, 62, 767–785. doi: 10.1111/j.1540–4560.2006.00486.x DuBois, D. L., Felner, R. D., Brand, S., Adan, A. M., & Evans, E. G. (1992). A prospective study of life stress, social support and adaptation in early adolescence. Child Development, 63, 542–557. doi: 10.2307/1131345 DuBois, D. L., Holloway, B. E., Valentine, J. C., & Cooper, H. (2002). Effectiveness of mentoring programs for youth: A meta-analytic review. American Journal of Community Psychology, 30, 157–197. doi: 10.1023/A:1014628810714 DuBois, D. L., Portillo, N., Rhodes, J. E., Silverthorn, N., & Valentine, J. C. (2011). How effective are mentoring programs for youth? A systematic assessment of the evidence. Psychological Science in the Public Interest, 12, 57–91. doi: 10.1177/1529100611414806 Dunham, P. J., Hurshman, A., Litwin, E., Gusella, J., Ellsworth, C., & Dodd, P. W. D. (1998). Computer-mediated social support: Single young mothers as a model system. American Journal of Community Psychology, 26, 281–306. doi: 10.1023/A:1022132720104 Eley, T. C. (2001). Contributions of behavioral genetic research: Quantifying genetic, shared environmental and nonshared environmental influences. In M. W. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of anxiety (pp. 45–59). London, UK: Oxford University Press. Ellison, N. B., Steinfield, C., & Lampe, C. (2007). The benefits of Facebook “friends:” Social capital and college students' use of online social network sites. Journal of ComputerMediated Communication, 12, 1143–1168. doi: 10.1111/j.1083–6101.2007.00367.x El-Sheikh, M., & Erath, S. A. (2011). Family conflict, autonomic nervous system functioning, and child adaptation: State of the science and future directions. Development and Psychopathology, 23, 703–721. doi: 10.1017/S0954579411000034 Evans, G. W., & Kim, P. (2007). Childhood poverty and health: Cumulative risk exposure and stress dysregulation. Psychological Science, 18, 953–957. doi: 10.1111/j.1467– 9280.2007.02008.x Evans, G. W., & Kim, P. (2013). Childhood poverty, chronic stress, self-regulation, and coping. Child Development Perspectives, 7, 43–48. doi: 10.1111/cdep.12013 Fantuzzo, J., Manz, P., Atkins, M., & Meyers, R. (2005). Peer-mediated treatment of socially withdrawn maltreated preschool children: Cultivating natural community resources. Journal of Clinical Child and Adolescent Psychology, 34, 320–325. doi: 10.1207/s15374424jccp3402_11

Fantuzzo, J., Sutton-Smith, B., Atkins, M., Myers, R., Stevenson, H., Coolahan, K.,… Manz, P. (1996). Community-based resilient peer treatment of withdrawn maltreated preschool children. Journal of Consulting and Clinical Psychology, 64, 1377–1386. doi: 10.1037/0022– 006X.64.6.1377 Fischer, C. S. (1982). To dwell among friends: Personal networks in town and city. Chicago, IL: University of Chicago Press. Fisher, J. D., Nadler, A., & Witcher-Alagna, S. (1982). Recipient reactions to aid. Psychological Bulletin, 91, 27–54. doi: 10.1037/0033–2909.91.1.27 Fisher, P. A., Burraston, B., & Pears, K. (2005). The Early Intervention Foster Care Program: Permanent placement outcomes from a randomized trial. Child Maltreatment, 10, 61–71. doi: 10.1177/1077559504271561 Fisher, P. A., & Kim, H. K. (2007). Intervention effects on foster preschoolers' attachmentrelated behaviors from a randomized trial. Prevention Science, 8, 161–170. doi: 10.1007/s11121–007–0066–5 Fisher, P. A., & Stoolmiller, M. (2008). Intervention effects on foster parent stress: Associations with child cortisol levels. Development and Psychopathology, 20, 1003–1021. doi: 10.1017/S0954579408000473 Fisher, P. A., Stoolmiller, M., Gunnar, M. R., & Burraston, B. O. (2007). Effects of a therapeutic intervention for foster preschoolers on diurnal cortisol activity. Psychoneuroendocrinology, 32, 892–905. doi: 10.1016/j.psyneuen.2007.06.008 Fisher, P. A., Van Ryzin, M. J., & Gunnar, M. R. (2011). Mitigating HPA axis dysregulation associated with placement changes in foster care. Psychoneuroendocrinology, 36, 531–539. doi: 10.1016/j.psyneuen.2010.08.007 Freisthler, B., Holmes, M. H., & Wolf, J. P. (2014). The dark side of social support: Understanding the role of social support, drinking behaviors and alcohol outlets for child physical abuse. Child Abuse & Neglect. 38, 1106–1119. doi: 10.1016/j.chiabu.2014.03.011 Froland, C., Pancoast, D. L., Chapman, N. J., & Kimboko, P. J. (1981). Linking formal and informal support systems. In B. H. Gottlieb (Ed.), Social networks and social support (pp. 259–275). Beverly Hills, CA: Sage. Fuligni, A. J., & Pedersen, S. (2002). Family obligation and the transition to young adulthood. Developmental Psychology, 38, 856–868. doi: 10.1037/0012–1649.38.5.856 Fuligni, A. J., Tseng, V., & Lam, M. (1999). Attitudes toward family obligations among American adolescents with Asian, Latin American, and European backgrounds. Child Development, 70, 1030–1044. doi: 10.1111/1467–8624.00075 Fuligni, A. J., Yip, T., & Tseng, V. (2002). The impact of family obligation on the daily

activities and psychological well-being of Chinese American adolescents. Child Development, 73, 302–314. doi: 10.1111/1467–8624.00407 Furman, W., & Buhrmester, D. (1985). Children's perceptions of the personal relationships in their social networks. Developmental Psychology, 21, 1016–1024. doi: 10.1037/0012– 1649.21.6.1016 Furman, W., & Buhrmester, D. (1992). Age and sex differences in perceptions of networks of personal relationships. Child Development, 63, 103–115. doi: 10.2307/1130905 Garbarino, J., & Kostelny, K. (1992). Child maltreatment as a community problem. Child Abuse & Neglect, 16, 455–464. doi: 10.1016/0145–2134(92)90062-V Garbarino, J., & Sherman, D. (1980a). High-risk neighborhoods and high-risk families: The human ecology of child maltreatment. Child Development, 51, 188–198. doi: 10.2307/1129606 Garbarino, J., & Sherman, D. (1980b). Identifying high-risk neighborhoods. In J. Garbarino & S. H. Stocking (Eds.), Protecting children from abuse and neglect (pp. 94–108). San Francisco, CA: Jossey-Bass. Gaudin, J. M., & Polansky, N. A. (1986). Social distancing of the neglectful family: Sex, race, and social class influences. Child and Youth Services Review, 8, 1–12. doi: 10.1016/0190– 7409(86)90022–8 Gaudin, J. M., Wodarski, J. S., Arkinson, M. K., & Avery, L. S. (1990–1991). Remedying child neglect: Effectiveness of social network interventions. Journal of Applied Social Sciences, 15, 97–123. Gazelle, H., & Ladd, G. W. (2003). Anxious solitude and peer exclusion: A diathesis–stress model of internalizing trajectories in childhood. Child Development, 74, 257–278. doi: 10.1111/1467–8624.00534 General Accounting Office. (1990). Home visiting: A promising early intervention strategy for at-risk families (GAO/HRD-90–83). Washington, DC: Government Printing Office. Gerlsma, C., Emmelkamp, P. M. G., & Arrindell, W. A. (1990). Anxiety, depression, and perception of early parenting: A meta-analysis. Clinical Psychology Review, 10, 251–277. doi: 10.1016/0272–7358(90)90062-F. Gomby, D. S., Culross, P. L., & Behrman, R. E. (1999). Home visiting: Recent program evaluations – Analysis and recommendations. Future of Children, 9, 4–26. Goodvin, R. (2004, June). High-risk mothers' experiences with Early Head Start: A qualitative exploration. Paper presented at the National Head Start Research Conference, Washington, DC. Gottlieb, B. (1983). Social support strategies. Beverly Hills, CA: Sage.

Gottlieb, B. (1985). Theory into practice: Issues that surface in planning interventions which mobilize support. In I. G. Sarason & B. R. Sarason (Eds.), Social support: Theory, research and applications (pp. 417–437). The Hague, the Netherlands: Martinus Nijhoff. Gottlieb, B. H. (2000). Selecting and planning support interventions. In S. Cohen, L. G. Underwood, & B. H. Gottlieb (Eds.), Social support measurement and intervention (pp. 195– 220). New York, NY: Oxford University Press. Gottlieb, B. H., & Bergen, A. E. (2010). Social support concepts and measures. Journal of Psychosomatic Research, 69, 511–520. doi: 10.1016/j.jpsychores.2009.10.001 Gottman, J. M., & Parker, J. G. (Eds.) (1987). Conversations of friends: Speculations on affective development. New York, NY: Cambridge University Press. Gray, M. (2009). Out in the country: Youth, media, and queer visibility in rural America. New York, NY: New York University Press. Greenberg, M. S., & Westcott, D. R. (1983). Indebtedness as a mediator of reactions to aid. In J. D. Fisher, A. Nadler, & B. M. DePaulo (Eds.), New directions in helping, Vol. 1. Recipient reactions to aid (pp. 85–112). New York, NY: Academic Press. Greenfield, P. M., Keller, H., Fuligni, A., & Maynard, A. (2003). Cultural pathways through universal development. Annual Review of Psychology, 54, 461–490. doi: 10.1146/annurev.psych.54.101601.145221 Gross, E. F. (2009). Logging on, bouncing back: An experimental investigation of online communication following social exclusion. Developmental Psychology, 45, 1787–1793. doi: 10.1037/a0016541 Grossman, J. B., & Tierney, J. P. (1998). Does mentoring work? An impact study of the Big Brothers Big Sisters program. Evaluation Review, 22, 403–426. doi: 10.1177/0193841X9802200304 Grusec, J. E., & Goodnow, J. J. (1994). Impact of parental discipline methods on the child's internalization of values: A reconceptualization of current points of view. Developmental Psychology, 30, 4–19. doi: 10.1111/j.1467–8624.2010.01426.x Guevremont, D. C., MacMillan, V. M., Shawchuck, C. R., & Hansen, D. J. (1989). A peermediated intervention with clinic-referred socially isolated girls: Generalization, maintenance, and social validation. Behavior Modification, 13, 32–50. doi: 10.1177/01454455890131002 Gunnar, M. R., & Donzella, B. (2002). Social regulation of the cortisol levels in early human development. Psychoneuroendocrinology, 27, 199–220. doi: 10.1016/S0306– 4530(01)00045–2 Gunnar, M. R., Morison, S. J., Chisholm, K., & Schuder, M. (2001). Salivary cortisol levels in children adopted from Romanian orphanages. Development and Psychopathology, 13, 611–

628. doi: 10.1017/S095457940100311X Gunnarsson, L., & Cochran, M. (1990). The support networks of single parents: Sweden and the United States. In M. Cochran, M. Larner, D. Riley, L. Gunnarsson, & C. R. Henderson (Eds.), Extending families: The social networks of parents and their children (pp. 105–116). New York, NY: Cambridge University Press. Haber, M. G., Cohen, J. L., Lucas, T., & Baltes, B. B. (2007). The relationship between selfreported received and perceived social support: A meta-analytic review. American Journal of Community Psychology, 39, 133–144. doi: 10.1007/s10464–007–9100–9 Halpern, R. (2000). Early childhood intervention for low income children and families. In J. P. Shonkoff & S. J. Meisels (Eds.), Handbook of early childhood intervention (2nd ed., pp. 361–386). New York, NY: Cambridge University Press. Hartup, W. W., & Stevens, N. (1999). Friendships and adaptation across the life span. Current Directions in Psychological Science, 8, 76–79. doi: 10.1111/1467–8721.00018 Hazzard, A., Celanno, M., Collins, M., & Markov, Y. (2002). Effects of STARBRIGHT World on knowledge, social support, and coping in hospitalized children with sickle cell disease and asthma. Children's Health, 31, 69–86. doi: 10.1207/S15326888CHC3101_5 Heller, K., & Rook, K. S. (2001). Distinguishing the theoretical functions of social ties: Implications for support interventions. In B. R. Sarason & S. Duck (Eds.), Personal relationships: Implications for clinical and community psychology (pp. 119–139). Chichester, UK: Wiley. Heller, K., & Swindle, R. W. (1983). Social networks, perceived social support, and coping with stress. In R. D. Felner, L. A. Jason, J. N. Moritsugu, & S. S. Farber (Eds.), Preventive psychology: Theory, research, and practice (pp. 87–103). New York, NY: Pergamon Press. Hettema, J. M., Neale, M. C., & Kendler, K. S. (2001). A review and meta-analysis of the genetic epidemiology of anxiety disorders. American Journal of Psychiatry, 158, 1568–1578. doi: 10.1176/appi.ajp.158.10.1568 Hibel, L., Granger, D. A., Blair, C., Cox, M. J., & the Family Life Project Key Investigators (2011). Maternal sensitivity buffers the adrenocortical implications of intimate partner violence exposure during early childhood. Development and Psychopathology, 23, 689–701. doi: 10.1017/S0954579411000010 Hirsch, B. J. (2005). A place to call home: After-school programs for urban youth. Washington, DC: American Psychological Association. Hirshfeld, D. R., Biederman, J., Brody, L., Faraone, S. V., & Rosenbaum, J. F. (1997a). Associations between expressed emotion and child behavioral inhibition and psychopathology: A pilot study. Journal of the American Academy of Child and Adolescent Psychiatry, 36, 205–213. doi: 10.1097/00004583–199702000–00011

Hirshfeld, D. R., Biederman, J., Brody, L., Faraone, S. V., & Rosenbaum, J. F. (1997b). Expressed emotion toward children with behavioral inhibition: Associations with maternal anxiety disorder. Journal of the American Academy of Child and Adolescent Psychiatry, 36, 910–917. doi: 10.1097/00004583–199707000–00012 Homel, R., Burns, A., & Goodnow, J. (1987). Parental social networks and child development. Journal of Social and Personal Relationships, 4, 159–177. doi: 10.1177/0265407587042004 Hooley, J. M., Orley, J., & Teasdale, J. D. (1986). Levels of expressed emotion and relapse in depressed patients. British Journal of Psychiatry, 148, 642–647. doi: 10.1192/bjp.148.6.642 Hooley, J. M., & Richters, J. E. (1995). Expressed emotion: A developmental perspective. In D. Cicchetti & S. Toth (Eds). Emotion, cognition, and representation. Rochester, NY: University of Rochester Press. Hostinar, C. E., Sullivan, R. M., & Gunnar, M. R. (2014). Psychobiological mechanisms underlying the social buffering of the hypothalamic-pituitary-adrenocortical axis: A review of animal models and human studies across development. Psychological Bulletin, 140, 256–282. doi: 10.1037/a0032671 House, J. S., & Kahn, R. L. (1985). Measures and concepts of social support. In J. Eckenrode (Ed.), The social context of coping (pp. 213–237). New York, NY: Plenum Press. House, J. H., Umberson, D., & Landis, K. R. (1988). Structures and processes of social support. In W. R. Scott & J. Blake (Eds.), Annual review of sociology (Vol. 14, pp. 293–318). Palo Alto, CA: Annual Reviews. doi: 10.1146/annurev.so.14.080188.001453 Hrdy, S. B. (2009). Mothers and others. Cambridge, MA: Harvard University Press. Hurd, N. M., Stoddard, S. A., & Zimmerman, M. A. (2013). Neighborhoods, social support, and African American adolescents' mental health outcomes: A multilevel path analysis. Child Development, 84, 858–874. doi: 10.1111/cdev.12018 Jackson, Y., & Warren, J. S. (2000). Appraisal, social support, and life events: Predicting outcome behavior in school-age children. Child Development, 71, 1441–1457. doi: 10.1111/1467–8624.00238 Jacobson, D. (1987). The cultural context of social support and social networks. Medical Anthropology Quarterly, 1, 42–67. doi: 10.1525/maq.1987.1.1.02a00030 Jacobson, S. W., & Frye, K. F. (1991). Effect of maternal social support on attachment: Experimental evidence. Child Development, 62, 572–582. doi: 10.2307/1131132 Jennings, K. D., Stagg, V., & Connors, R. E. (1991). Social networks and mothers' interactions with their preschool children. Child Development, 62, 966–978. doi: 10.2307/1131146 Jiang, L, C., Bazarova, N. N., & Hancock, J. T. (2011). The disclosure-intimacy link in computer-mediated communication: An attributional extension of the hyperpersonal model.

Human Communication Research, 37, 58–77. doi: 10.1111/j.1468–2958.2010.01393.x Johnston, K., & Brinamen, C. (2006). Mental health consultation in child care. Washington, DC: Zero to Three Press. Kagitcibasi, C. (2005). Autonomy and relatedness in cultural context: Implications for self and family. Journal of Cross-Cultural Psychology, 36, 403–422. doi: 10.1177/0022022105275959 Kaufman, J., Yang, B.-Z., Douglas-Palumberi, H., Houshyar, S., Lipschitz, D., Krystal, J. H., & Gelernter, J. (2004). Social supports and serotonin transporter gene moderate depression in maltreated children. Proceedings of the National Academy of Sciences, 101, 17316–17321. doi: 10.1073/pnas.0404376101 Kaul, M., & Lackey, B. (2003). Where is the support in perceived support? The role of generic relationship satisfaction and enacted support in perceived support's relation to low distress. Journal of Social and Clinical Psychology, 22, 59–78. doi: 10.1521/jscp.22.1.59.22761 Keller, H. (2007). Cultures of infancy. Mahwah, NJ: Erlbaum. Kiecolt-Glaser, J. K., Gouin, J.-P., & Hantsoo, L. (2010). Close relationships, inflammation, and health. Neuroscience and Biobehavioral Reviews, 35, 33–38. doi: 10.1016/j.neubiorev.2009.09.003 Kochanska, G. (2002). Mutually responsive orientation between mothers and their young children: A context for the early development of conscience. Current Directions in Psychological Science, 11, 191–195. doi: 10.1111/1467–8721.00198 Korbin, J. E. (1989). Fatal maltreatment by mothers: A proposed framework. Child Abuse & Neglect, 13, 481–489. doi: 10.1016/0145–2134(89)90052–5 Korbin, J. E. (1991, November). “Good mothers,” “babykillers,” and fatal child abuse. Paper presented to the annual meeting of the American Anthropological Association, Chicago. Korbin, J. E. (1995). Social networks and family violence in cross-cultural perspective. In G. B. Melton (Ed.), Nebraska symposium on motivation, Vol. 42. The individual, the family, and social good: Personal fulfillment in times of change (pp. 107–134). Lincoln, NE: University of Nebraska Press. Ladd, G. W., Hart, C. H., Wadsworth, E. M., & Golter, B. S. (1988). Preschoolers' peer networks in nonschool settings: Relationship to family characteristics and school adjustment. In S. Salzinger, J. Antrobus, & M. Hammer (Eds.), Social networks of children, adolescents, and college students (pp. 61–92). Hillsdale, NJ: Erlbaum. Ladd, G. W., & Le Sieur, K. D. (1994). Parents and children's peer relationships. In M. H. Bornstein (Ed.), Handbook of parenting, Vol. 4. Applied and practical parenting (pp. 377– 409). Hillsdale, NJ: Erlbaum.

Laible, D. J., & Thompson, R. A. (2007). Early socialization: A relational perspective. In J. Grusec & P. Hastings (Eds.), Handbook of socialization (pp. 181–207). New York, NY: Guilford Press. Lakey, B., & Lutz, C. J. (1996). Social support and preventive and therapeutic interventions. In G. R. Pierce, B. R. Sarason, & I. G. Sarason (Eds.), Handbook of social support and the family (pp. 435–465). New York, NY: Plenum Press. Langhout, R. D., Rhodes, J. E., & Osborne, L. N. (2004). An exploratory study of youth mentoring in an urban context: Adolescents' perceptions of relationship styles. Journal of Youth and Adolescence, 33, 293–306. doi: 10.1023/B:JOYO.0000032638.85483.44 Lansford, J. E., Chang, L., Dodge, K. A., Malone, P. S., Oburu, P., Palmerus, K.,… Quinn, N. (2005). Physical discipline and children's adjustment: Cultural normativeness as a moderator. Child Development, 76, 1234–1246. doi: 10.1111/j.1467–8624.2005.00847.x Larner, M. (1990a). Changes in network resources and relationships over time. In M. Cochran, M. Larner, D. Riley, L. Gunnarsson, & C. R. Henderson (Eds.), Extending families: The social networks of parents and their children (pp. 181–204). New York, NY: Cambridge University Press. Larner, M. (1990b). Local residential mobility and its effects on social networks: A crosscultural comparison. In M. Cochran, M. Larner, D. Riley, L. Gunnarsson, & C. R. Henderson (Eds.), Extending families: The social networks of parents and their children (pp. 205–229). New York, NY: Cambridge University Press. Lewis, M. W., & Lewis, A. C. (1996). Peer helping programs. Helper role, supervisor training, and suicidal behavior. Journal of Counseling and Development, 74, 307–314. doi: 10.1002/j.1556–6676.1996.tb01871.x Lieberman, A. F., Ghosh Ippen, C., & Van Horn, P. (2006). Child-parent psychotherapy: 6month follow-up of a randomized controlled trial. Journal of the American Academy of Child and Adolescent Psychiatry, 45, 913–918. doi: 10.1097/01.chi.0000222784.03735.92 Lieberman, A. F., & Van Horn, P. (2008). Psychotherapy with infants and young children: Repairing the effects of stress and trauma on early attachment. New York, NY: Guilford Press Lieberman, A. F., Van Horn, P., & Ghosh Ippen, C. (2005). Toward evidence-based treatment: Child-Parent Psychotherapy with preschoolers exposed to marital violence. Journal of the American Academy of Child and Adolescent Psychiatry, 44, 1241–1248. doi: 10.1097/01.chi.0000181047.59702.58 Litwak, E., & Szelenyi, I. (1969). Primary group structures and their functions: Kin, neighborhoods, and friends. American Sociological Review, 34, 465–481. Lovell, M. L., & Hawkins, J. D. (1988). An evaluation of a group intervention to increase the

personal social networks of abusive mothers. Children and Youth Services Review, 10, 175– 188. doi: 10.1016/0190–7409(88)90001–1. Luby, J. L., Barch, D. M., Belden, A., Gaffrey, M. S., Tillman, R., Babb, C.,… Botteron, K. N. (2012). Maternal support in early childhood predicts larger hippocampal volumes at school age. Proceedings of the National Academy of Sciences, 109, 2854–2859. doi: 10.1073/pnas.1118003109 Lupien, S. J., McEwen, B. S., Gunnar, M. R., & Heim, C. (2009). Effects of stress throughout the lifespan on the brain, behaviour and cognition. Nature Reviews Neuroscience, 10, 434– 445. doi: 10.1038/nrn2639 Lynch, M., & Cicchetti, D. (2002). Links between community violence and the family system: Evidence from children's feelings of relatedness and perceptions of parent behavior. Family Process, 41, 519–532. doi: 10.1111/j.1545–5300.2002.41314.x Lyons-Ruth, K., & Jacobvitz, D. (2008). Attachment disorganization: Genetic factors, parenting contexts, and developmental transformation from infancy to adulthood. In J. Cassidy & P. R. Shaver (Eds.), Handbook of attachment (2nd ed., pp. 666–697). New York, NY: Guilford Press Maccoby, E. E. (1984). Socialization and developmental change. Child Development, 55, 317–328. doi: 10.2307/1129945 MacPhee, D., Fritz, J., & Miller-Heyl, J. (1996). Ethnic variations in personal social networks and parenting. Child Development, 67, 3278–3295. doi: 10.2307/1131779 Margie, N. G., & Phillips, D. (1999). Revisiting home visiting: Summary of a workshop. Washington, DC: National Academy of Sciences Press. Markus, H. R., & Kitayama, S. (1991). Culture and the self: Implications for cognition, emotion, and motivation. Psychological Review, 98, 224–253. doi: 10.1037/0033– 295X.98.2.224 McKenna, K. Y., & Bargh, J. A. (1998). Coming out in the age of the Internet: Identity “demarginalization” through virtual group participation. Journal of Personality and Social Psychology, 75, 681–694. doi: 10.1037/0022–3514.75.3.681 Meaney, M. J. (2010). Epigenetics and the biological definition of gene × environment interactions. Child Development, 81, 41–79. doi: 10.1111/j.1467–8624.2009.01381.x Medrich, E. A., Roizen, J. A., Rubin, V., & Buckley, S. (1982). The serious business of growing up: A study of children's lives outside school. Berkeley, CA: University of California Press. Meehan, B. T., Hughes, J. N., & Cavell, T. A. (2003). Teacher-student relationships as compensatory resources for aggressive children. Child Development, 74, 1145–1157. doi:

10.1111/1467–8624.00598 Melson, G. F., Ladd, G. W., & Hsu, H.-C. (1993). Maternal support networks, maternal cognitions, and young children's social and cognitive development. Child Development, 64, 1401–1417. doi: 10.2307/1131542 MENTOR/National Mentoring Partnership. (2006). Mentoring in America 2005: A snapshot of the current state of mentoring. Boston, MA: National Mentoring Partnership. Miller, G. E., Chen, E., & Parker, K. J. (2011). Psychological stress in childhood and susceptibility to the chronic diseases of aging: Moving toward a model of behavioral and biological mechanisms. Psychological Bulletin, 137, 959–997. doi: 10.1037/a0024768. Miller, J. L., & Whittaker, J. K. (1988). Social services and social support: Blended programs for families at risk for child maltreatment. Child Welfare, 67, 161–174. Nachmias, M., Gunnar, M., Mangelsdorf, S., Parritz, R. H., & Buss, K. (1996). Behavioral inhibition and stress reactivity: The moderating role of attachment security. Child Development, 67, 508–522. doi: 10.1111/j.1467–8624.1996.tb01748.x Nangle, D. W., Erdley, C. A., Newman, J. E., Mason, C. A., & Carpenter, E. M. (2003). Popularity, friendship quantity, and friendship quality: Interactive influences of children's loneliness and depression. Journal of Clinical Child and Adolescent Psychology, 32, 546– 555. doi: 10.1207/S15374424JCCP3204_7 Naumova, O. U., Lee, M., Koposov, R., Szyf, M., Dozier, M., & Grigorenko, E. L. (2012). Differential patterns of whole-genome DNA methylation in institutionalized children and children raised by their biological parents. Development and Psychopathology, 24, 143–155. doi: 10.1017/S0954579411000605 Nemoto, T. U. (1998). Subjective norms toward social support among Japanese American elderly in New York City: Why help does not always help. Journal of Community Psychology, 26, 293–316. doi: 10.1002/(SICI)1520–6629(199807)26:4 Newcomb, A. F., & Bagwell, C. L. (1995). Children's friendship relations: A meta-analytic review. Psychological Bulletin, 117, 306–347. doi: 10.1037/0033–2909.117.2.306 Oberlander, T. F., Weinberg, J., Papsdorf, M., Grunau, R., Misri, S., & Devlin, A. M. (2008). Prenatal exposure to maternal depression, neonatal methylation of human glucocorticoid receptor gene (NR3C1) and infant cortisol stress responses. Epigenetics, 3, 97–106. doi: 10.4161/epi.3.2.6034 Odgers, C. L., Moffitt, T. E., Tach, L. M., Sampson, R. J., Taylor, A., Matthews, C. L., & Caspi, A. (2009). The protective effects of neighborhood collective efficacy on British children growing up in deprivation: A developmental analysis. Developmental Psychology, 45, 942–957. doi: 10.1037/a0016162

O'Donnell, L., & Steuve, A. (1983). Mothers as social agents: Structuring the community activities of school-age children. In H. Lopata & J. H. Pleck (Eds.), Research on the interweave of social roles: Jobs and families, Vol. 3. Families and jobs (pp. 113–129). Greenwich, CT: JAI. Oldenburg, C. M., & Kerns, K. A. (1997). Associations between peer relationships and depressive symptoms: Testing moderator effects of gender and age. Journal of Early Adolescence, 17, 319–337. doi: 10.1177/0272431697017003004 Olds, D. L. (2006). The nurse–family partnership: An evidence-based preventive intervention. Infant Mental Health Journal, 27, 5–25. doi: 10.1002/imhj.20077 Parke, R. D., & Bhavnagri, N. P. (1989). Parents as managers of children's peer relationships. In D. Belle (Ed.), Children's social networks and social supports (pp. 241–259). New York, NY: Wiley. Parke, R. D., & Buriel, R. (2006). Socialization in the family: Ethnic and ecological perspectives. In W. Damon & R. M. Lerner (Eds.), Handbook of child psychology, Vol. 3. Social, emotional, and personality development (N. Eisenberg, Vol. Ed.), 6th ed. (pp. 429– 504). Hoboken, NJ: Wiley. Parker, J. G., & Asher, S. R. (1993). Friendship and friendship quality in middle childhood: Links with peer group acceptance and feelings of loneliness and social dissatisfaction Developmental Psychology, 29, 611–621. doi: 10.1037/0012–1649.29.4.611. Parker, J. G., Rubin, K. H., Erath, S. A., Wojslawowicz, J., & Buskirk, A. A. (2006). Peer relationships, child development, and adjustment: A developmental psychopathology perspective. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, 2nd Ed., Vol. 1 Theory and method (pp. 419–493). Hoboken, NJ: New York, NY: Wiley. Pavuluri, M. N. (2008). What works for bipolar kids: Help and hope for parents. New York, NY: Guilford Press Pavuluri, M. N., Graczyk, P. A., Henry, D. B., Carbray, J. A., Heidenreich, J., & Miklowitz, D. J. (2004). Child-focused cognitive-behavioral therapy for pediatric bipolar disorder: Development and preliminary results. Journal of the American Academy of Child and Adolescent Psychiatry, 43, 528–537. doi: 10.1097/00004583–200405000–00006 Pelton, L. H. (1989). For reasons of poverty: A critical analysis of the public child welfare system in the United States. New York, NY: Praeger. Pelton, L. H. (1994). The role of material factors in child abuse and neglect. In G. B. Melton & F. D. Barry (Eds.), Protecting children from abuse and neglect: Foundation for a new national strategy (pp. 131–181). New York, NY: Guilford Press. Polansky, N. A., Ammons, P. W., & Gaudin, J. M. (1985). Loneliness and isolation in child neglect. Social Casework, 66, 38–47.

Polansky, N. A., Chalmers, M. A., Buttenwieser, E., & Williams, D. P. (1981). Damaged parents: An anatomy of child neglect. Chicago, IL: University of Chicago Press. Polansky, N. A., & Gaudin, J. M. (1983). Social distancing of the neglectful family. Social Service Review, 57, 196–208. Polansky, N. A., Gaudin, J. M., Ammons, P. W., & David, K. B. (1985). The psychological ecology of the neglectful mother. Child Abuse & Neglect, 9, 265–275. doi: 10.1016/0145– 2134(85)90019–5 Pomerantz, E. M., & Dong, W. (2006). The effects of mothers' perceptions of children's competence: The moderating role of mothers' theories of competence. Developmental Psychology, 42, 950–961. doi: 10.1037/0012–1649.42.5.950 Pomerantz, E. M., Moorman, E. A., & Litwack, S. D. (2007). The how, whom, and why of parents' involvement in children's academic lives: More is not always better. Review of Educational Research, 77, 373–410. doi: 10.3102/003465430305567 Powell, D. R. (1979). Family-environment relations and early childrearing: The role of social networks and neighborhoods. Journal of Research and Development in Education, 13, 1–11. Ranney, J. D., & Troop-Gordon, W. (2012). Computer-mediated communication with distant friends: Relations with adjustment during students first semester in college. Journal of Educational Psychology, 104, 848–861. doi: 10.1037/a0027698 Reid, M., Landesman, S., Treder, R., & Jaccard, J. (1989). “My family and friends”: Six- to twelve-year-old children's perceptions of social support. Child Development, 60, 896–910. doi: 10.2307/1131031 Repetti, R. L., Robles, T. F., & Reynolds, B. (2011). Allostatic processes in the family. Development and Psychopathology, 23, 921–938. doi: 10.1017/S095457941100040X Rhodes, G. L., & Lakey, B. (1999). Social support and psychological disorder: Insights from social psychology. In R. M. Kowalsky & M. R. Leary (Eds.), The social psychology of emotional and behavioral problems (pp. 281–309). Washington, DC: American Psychological Association. Rhodes, J. E. (2002). Stand by me: The risks and rewards of mentoring today's youth. Cambridge, MA: Harvard University Press. Rhodes, J. E. (2005). A model of youth mentoring. In D. L. DuBois & M. J. Karcher (Eds.), Handbook of youth mentoring (pp. 30–43). Thousand Oaks, CA: Sage. Rhodes, J. E., & DuBois, D. L. (2008). Mentoring relationships and programs for youth. Current Directions in Psychological Science, 17, 254–258. doi: 10.1111/j.1467– 8721.2008.00585.x. Rhodes, J. E., Ebert, L., & Fischer, K. (1992). Natural mentors: An overlooked resource in the

social networks of young, African American mothers. American Journal of Community Psychology, 20, 445–461. doi: 10.1007/BF00937754 Rhodes, J. E., Reddy, R., & Grossman, J. B. (2005). The protective influence of mentoring on adolescents' substance use: Direct and indirect pathways. Applied Developmental Science, 9, 31–47. doi: 10.1207/s1532480xads0901_4 Robinson, N. S., & Garber, J. (1995). Social support and psychopathology across the life span. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, Vol. 2. Risk, disorder, and adaptation (pp. 162–209). New York, NY: Wiley. Rodriguez, N., Bingham Mira, C., Myers, H. F., Morris, J. K., & Cardoza, D. (2003). Family or friends: Who plays a greater supportive role for Latino college students? Cultural Diversity and Ethnic Minority Psychology, 9, 236–250. doi: 10.1037/1099–9809.9.3.236 Rogoff, B. (2003). The cultural nature of human development. New York, NY: Oxford University Press. Rogosch, F. A., Cicchetti, D., & Toth, S. L. (2004). Expressed emotion in multiple subsystems of the families of toddlers with depressed mothers. Development and Psychopathology, 16, 689–709. doi: 10.1017/S0954579404004730. Rose, A. J., Carlson, W., & Waller, E. M. (2007). Prospective associations of co-rumination with friendship and emotional adjustment: Considering the socioemotional trade-offs of corumination. Developmental Psychology, 43, 1019–1031. doi: 10.1037/0012–1649.43.4.1019 Rubin, K. H., Coplan, R., Chen, X., Bowker, J., & McDonald, K. L. (2011). Peer relationships in childhood. In M. H. Bornstein & M. E. Lamb (Eds.), Developmental science: An advanced textbook (6th ed., pp. 519–570). New York, NY: Psychology Press. Sagobal, F., Marin, G., Otero-Sabogal, R., Van Oss Marin, B., & Perez Stable, E. J. (1987). Hispanic familialism and acculturation: What changes and what doesn't? Hispanic Journal of Behavioral Sciences, 9, 397–412. doi: 10.1177/07399863870094003. Salzinger, S., Antrobus, J., & Hammer, M. (Eds.) (1988). Social networks of children, adolescents, and college students. Hillsdale, NJ: Erlbaum. Sanchez, B., & Reyes, O. (1999). Descriptive profile of the mentorship relationships of Latino adolescents. Journal of Community Psychology, 27, 299–302. doi: 10.1002/(SICI)1520– 6629(199905)27:3 Sarason, B. R., Shearin, E. N., Pierce, G. R., & Sarason, I. G. (1987). Interrelations of social support measures: Theoretical and practical implications. Journal of Personality and Social Psychology, 50, 845–855. doi: 10.1037/0022–3514.52.4.813 Seagull, E. A. W. (1987). Social support and child maltreatment: A review of the evidence. Child Abuse & Neglect, 11, 41–52. doi: 10.1016/0145–2134(87)90032–9

Shinn, M., Lehmann, S., & Wong, N. W. (1984). Social interaction and social support. Journal of Social Issues, 40, 55–76. doi: 10.1111/j.1540–4560.1984.tb01107.x Shumaker, S. A., & Brownell, A. (1984). Toward a theory of social support: Closing conceptual gaps. Journal of Social Issues, 40, 11–36. doi: 10.1111/j.1540– 4560.1984.tb01105.x Simkins, S., & Parke, R. (2002). Do friends and nonfriends behave differently? A social relations analysis of children's behavior. Merrill-Palmer Quarterly, 48, 263–283. Sinclair, J. J., Pettit, G. S., Harrist, A. W., Dodge, K. A., & Bates, J. E. (1994). Encounters with aggressive peers in early childhood: Frequency, age differences, and correlates of risk for behaviour problems. International Journal of Behavioural Development, 17, 675–696. doi: 10.1177/016502549401700407 Small, M. L. (2009). Unanticipated gains: Origins of network inequality in everyday life. Oxford, UK: Oxford University Press. Smetana, J. (2006). Social-cognitive domain theory: Consistencies and variations in children's moral and social judgments. In M. Killen & J. G. Smetana (Eds.), Handbook of moral development (pp. 119–154). Mahwah, NJ: Erlbaum. Snyder, J., Cramer, A., Afrank, J., & Patterson, G. R. (2005). The contributions of ineffective discipline and parental hostile attributions of child misbehavior to the development of conduct problems at home and school. Developmental Psychology, 41, 30–41. doi: 10.1037/0012– 1649.41.1.30 Spangler, G., & Grossmann, K. E. (1993). Biobehavioral organization in securely and insecurely attached infants. Child Development, 64, 1439–1450. doi: 10.2307/1131544 Strauss, M. A., Gelles, R. J., & Steinmetz, S. (1980). Behind closed doors: Violence in the American family. Garden City, NY: Doubleday/Anchor. Stronach, E. P., Toth, S. L., Rogosch, F., & Cicchetti, D. (2013). Preventive interventions and sustained attachment security in maltreated children. Development and Psychopathology, 25, 919–930. doi: 10.1017/S0954579413000278 Suomi, S. (2011). Risk, resilience, and gene-environment interplay in primates. Journal of the Canadian Academy of Child and Adolescent Psychiatry, 20, 289–298. Sweet, M. A., & Appelbaum, M. I. (2004). Is home visiting an effective strategy? A metaanalytic review of home visiting programs for families with young children. Child Development, 75, 1435–1456. doi: 10.1111/j.1467–8624.2004.00750.x Taylor, S. E. (2011). Social support: A review. In H. S. Friedman (Ed.), Oxford handbook of health psychology (pp. 189–214). New York, NY: Oxford University Press. Taylor, R. D., Casten, R., & Flickinger, S. M. (1993). Influence of kinship social support on the

parenting experiences and psychosocial adjustment of African-American adolescents. Developmental Psychology, 29, 382–388. doi: 10.1037/0012–1649.29.2.382 Thoits, P. A. (2011). Mechanisms linking social ties and support to physical and mental health. Journal of Health and Social Behavior, 52, 145–161. doi: 10.1177/0022146510395592 Thompson, R. A. (1995). Preventing child maltreatment through social support: A critical analysis. Thousand Oaks, CA: Sage. Thompson, R. A. (2001). Childhood anxiety disorders from the perspective of emotion regulation and attachment. In M. W. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of anxiety (pp. 160–182). London, UK: Oxford University Press. Thompson, R. A. (2006). The development of the person: Social understanding, relationships, self, conscience. In W. Damon & R. M. Lerner (Eds.), Handbook of child psychology (6th ed.), Vol. 3. Social, emotional, and personality development (N. Eisenberg, Vol. Ed.) (pp. 24– 98). Hoboken, NJ, New York, NY: Wiley. Thompson, R. A. (2008). Early attachment and later development: Familiar questions, new answers. In J. Cassidy & P. R. Shaver (Eds.), Handbook of attachment (2nd Ed.) (pp. 348– 365). New York, NY: Guilford Press Thompson, R. A. (2014a). Stress and child development. Future of Children, 25, 41–59. Thompson, R. A. (2014b). Why are relationships important to children's well-being? In A. Ben-Arieh, F. Casas, I. Frones, F. Cases, & J. E. Korbin (Eds.), Handbook of child well-being: Theories, methods, and policies in global perspective, Vol. 4 (pp. 1917–1954). Dordrecht, the Netherlands: Springer. Thompson, R. A. (2015a). Social support and child protection: Lessons learned and learning. Child Abuse & Neglect, 41(1), 19–29. doi: 10.1016/j.chiabu.2014.06.011 Thompson, R. A. (2015b). Relationships, regulation, and early development. In R. M. Lerner (Ed.), Handbook of child psychology and developmental science (7th ed.), Vol. 3: Social and emotional development (M. E. Lamb & C. García Coll, Vol. Eds.) (pp. pp. 201–246). Hoboken, NJ: Wiley. Thompson, R. A., Laible, D. J., & Robbennolt, J. K. (1997). Child care and preventing child maltreatment. In S. Reifel (Ed.), Advances in early education and day care, Vol. 9. Family policy and practice in early child care(C. J. Dunst & M. Wolery, Vol. Eds.) (pp. 173–202). Greenwich, CT: JAI Press. Thompson, R. A., & Ontai, L. (2000). Striving to do well what comes naturally: Social support, developmental psychopathology, and social policy. Development and Psychopathology, 12, 657–675. doi: 10.1017/S0954579400004065 Thompson, R. A., Scalora, M. J., Castrianno, L., & Limber, S. (1992). Grandparent visitation

rights: Emergent psychological and psycholegal issues. In D. K. Kagehiro & W. S. Laufer (Eds.), Handbook of psychology and law (pp. 292–317). New York, NY: Springer-Verlag. Toth, S. L., Maughan, A., Manly, J. T., Spagnola, M., & Cicchetti, D. (2002). The relative efficacy of two interventions in altering maltreated preschool children's representational models: Implications for attachment theory. Development and Psychopathology, 14, 877–908. doi: 10.1017.S095457940200411X Toth, S. L., Rogosch, F. A., Manly, J. T., & Cicchetti, D. (2006). The efficacy of toddler-parent psychotherapy to reorganize attachment in the young offspring of mothers with major depressive disorder: A randomized preventive trial. Journal of Consulting and Clinical Psychology, 74, 1006–1016. doi: 10.1037/0022–006X.74.6.1006 Toth, S. L., Rogosch, F. A., Sturge-Apple, M., & Cicchetti, D. (2009). Maternal depression, children's attachment security, and representational development: An organizational perspective. Child Development, 80, 192–208. doi: 10.1111/j.1467–8624.2008.01254.x Tracy, E. M., Whittaker, J. K., Boylan, F., Neitman, P., & Overstreet, E. (1995). Network interventions with high risk youth and families throughout the continuum of care. In I. Schwartz & P. AuClaire (Eds.), Home-based services for troubled children (pp. 122–167). Lincoln, NE: University of Nebraska Press. Triandis, H. C. (1989). The self and social behavior in differing cultural contexts. Psychological Review, 96, 506–520. doi: 10.1037/0033–295X.96.3.506 Uchino, B. N. (2004). Social support and physical health: Understanding the health consequences of relationships. New Haven, CT: Yale University Press. Ulrich-Lai, Y. M., & Herman, J. P. (2009). Neural regulation of endocrine and autonomic stress responses. Nature Reviews Neuroscience, 10, 397–409. doi: 10.1038/nrn2647 Umberson, D., Crosnoe, R., & Reczek, C. (2010). Social relationships and health behavior across the life course. Annual Review of Sociology, 36, 139–157. doi: 10.1146/annurev-soc070308–120011 Umberson, D., & Montez, J. K. (2010). Social relationships and health: A flashpoint for health policy. Journal of Health and Social Behavior, 51(Suppl.), S54–S66. doi: 10.1177/0022146510383501 Unger, D. G., & Powell, D. R. (1980). Supporting families under stress: The role of social networks. Family Relations, 29, 566–574. U.S. Advisory Board on Child Abuse and Neglect. (1991). Creating caring communities: Blueprint for an effective federal policy on child abuse and neglect. Washington, DC: Government Printing Office. U.S. Advisory Board on Child Abuse and Neglect. (1993). Neighbors helping neighbors: A

new national strategy for the protection of children. Washington, DC: Government Printing Office. Valkenburg, P. M., & Peter, J. (2007a). Online communication and adolescents' well-being: Testing the stimulation versus the displacement hypothesis. Journal of Computer Mediated Communication, 12, 292–315. doi: 10.1111/j.1083–6101.2007.00368.x Valkenburg, P. M., & Peter, J. (2007b). Preadolescents' and adolescents' online communication and their closeness to friends. Developmental Psychology, 43, 267–277. doi: 1037/0012– 1649.43.2.267.00368.x Valkenburg, P. M., & Peter, J. (2009a). The effects of instant messaging on the quality of adolescents' existing friendships: A longitudinal study. Journal of Communication, 59, 79–97. doi: 10.1111/j.1460–2466.2008.01405.x Valkenburg, P. M., & Peter, J. (2009b). Social consequences of the Internet for adolescents: A decade of research. Current Directions in Psychological Science, 18, 1–5. doi: 10.1111/j.1467–8721.2009.01595.x Vasey, M. W., & Dadds, M. R. (2001). An introduction to the developmental psychopathology of anxiety. In M. W. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of anxiety (pp. 1–22). London, UK: Oxford University Press. Vasey, M. W., & Ollendick, T. H. (2000). Anxiety. In M. Lewis & A. Sameroff (Eds.), Handbook of developmental psychopathology (2nd ed., pp. 511–529). New York, NY: Plenum Press. Vaux, A. (1985). Variations in social support associated with gender, ethnicity, and age. Journal of Social Issues, 41, 89–110. doi: 10.1111/j.1540–4560.1985.tb01118.x Vaux, A. (1988). Social support: Theory, research, and intervention. New York, NY: Praeger. Wagner, M. M., & Clayton, S. L. (1999). The Parents as Teachers Program: Results from two demonstrations. Future of Children, 9, 91–115. Wasik, B. H., & Bryant, D. M. (2000). Home visiting (2nd ed.). Newbury Park, CA: Sage. Waters, E., Kondo-Ikemura, K., Posada, G., & Richters, J. E. (1991). Learning to love: Mechanisms and milestones. In M. R. Gunnar & L. A. Sroufe (Eds.), Self processes and development. Minnesota Symposia on Child Psychology (Vol. 23, pp. 217–255). Hillsdale, NJ: Erlbaum. Weinraub, M., & Wolf, B. M. (1983). Effects of stress and social supports on mother-child interactions in single- and two-parent families. Child Development, 54, 1297–1311. doi: 10.2307/1129683 West, A. E., Henry, D. B., & Pavuluri, M. N. (2007). Maintenance model of integrated psychosocial treatment in pediatric bipolar disorder: A pilot feasibility study. Journal of the

American Academy of Child and Adolescent Psychiatry, 46, 205–212. doi: 10.1097/01.chi.0000246068.85577.d7 West, A. E., Jacobs, R. H., Westerholm, R., Lee, A., Carbray J., Heidenreich, J., & Pavuluri, M. N. (2009). Child and family-focused cognitive-behavioral therapy for pediatric bipolar disorder: Pilot study of group treatment format. Journal of the Canadian Academy of Child and Adolescent Psychiatry, 18, 239–246. Whitehill, J. M., Brockman, L. N., & Moreno, M. A. (2013). “Just talk to me”: Communicating with college students about depression disclosures on Facebook. Journal of Adolescent Health, 52, 122–127. doi: 10.1016/j.jadohealth.2012.09.015 Wortman, C. B., & Lehman, D. R. (1985). Reactions to victims of life crises: Support attempts that fail. In I. G. Sarason & B. R. Sarason (Eds.), Social support: Theory, research, and applications (pp. 463–489). Dordrecht, the Netherlands: Martinus Nijhoff. Young, L. (1964). Wednesday's children: A study of child neglect and abuse. New York, NY: McGraw-Hill. Zellman, G. (1990). Child abuse reporting and failure to report among mandated reporters. Journal of Interpersonal Violence, 5, 1–11. doi: 10.1177/088626090005001001 Zellman, G., & Anter, S. (1990). Mandated reporting and CPS: A study in frustration. Public Welfare, 48, 1–31. Zimmerman, M. A., Ramirez-Valles, J., Zapert, M., & Maton, K. I. (2000). A longitudinal study of stress-buffering effects for urban African-American male adolescent problem behaviors and mental health. Journal of Community Psychology, 28, 17–33. doi: 10.1002/(SICI)1520– 6629(200001)28:1

Chapter 4 Poverty and the Development of Psychopathology Martha E. Wadsworth, Gary W. Evans, Kathryn Grant, Jocelyn S. Carter, and Sophia Duffy RATIONALE AND BACKGROUND Why a Chapter on Poverty in this Handbook? Mental Health Costs of Poverty Ethical Issues and Social Justice U.S. Poverty Demographics Poverty Metrics Interrelations of Poverty and Psychopathology with Other Important Life Areas Direction of Causality POVERTY AND PSYCHOPATHOLOGY Internalizing Externalizing MEDIATORS AND MODERATORS OF POVERTY'S EFFECTS ON PSYCHOPATHOLOGY SES and Parenting SES and Stress SES, Executive Functioning, and Coping Brain and SES PRACTICE AND POLICY Community and/or Government Interventions: Policy Implications of Poverty and Psychology Research for Policy Individual and Family Interventions: Practice Implications of Poverty and Psychology Research for Practice Implications of Mediator Data for Future Interventions Implications of Moderator Data and Related Findings FUTURE DIRECTIONS REFERENCES

Rationale and Background Why a Chapter on Poverty in this Handbook? Affecting more than 16 million (22%) U.S. children, poverty places children at elevated risk for most forms of psychopathology. Of all of the factors that threaten the health and well-being of children and adolescents, poverty has the widest reach. It is a largely nonspecific risk factor, one that lays fertile ground for various genetic and nongenetic etiologic factors to emerge, take hold, and compound or interact with one another. Physical and mental health disparities stemming from socioeconomic inequality and poverty are well documented (Evans, Wolf, & Adler, 2012). Although the most startling effects, such as premature death, occur mostly during adulthood (Gruenewald, Cohen, Matthews, Tracy, & Seeman, 2012), research has clearly shown that the roots of health disparities take hold during childhood (e.g., Chen, Hanson, et al., 2006; Evans, Chen, Miller, & Seeman, 2012). Poor children are at significantly elevated risk for developing a wide array of internalizing and externalizing symptoms and disorders (Grant et al., 2003), including higher rates of anxiety, depression, and aggressive behavior (Miech, Caspi, Moffitt, Wright, & Silva, 1999; Wadsworth & Santiago, 2008), as well as other types of pathology as wide-ranging as schizophrenia and delinquency (McBride Murry, Berkel, Gaylord-Hardin, Copeland-Linder, & Nation, 2011). Poverty's pernicious effects are widespread for several reasons. First, poverty can have direct effects on children's health and development, stemming from problems such as hunger, exposure to toxins, and increased incidence of accidents and injuries (Evans, 2004). Poverty also increases a child's risk for exposure to a variety of known familial and parental risk factors, such as disrupted parenting, parental absence, interparental conflict, family violence, child maltreatment, and substance abuse, as well as parental psychopathology, such as depression, substance abuse, and antisocial behavior (Wadsworth et al., 2008). Additionally, poor children are often exposed to negative neighborhood characteristics, such as crime, deviant peer groups, poor-quality schools, and a variety of environmental hazards (Leventhal & Brooks-Gunn, 2000). An additional feature of poor children's lives is that these problems and exposures often co-occur, and Evans (2004; Evans & Kim, 2010) has shown that an accumulation of environmental and family risks takes the heaviest toll. Hence, poor children face numerous and varied risk factors in their homes, within their families, in their neighborhoods, and in the schools that make them more likely to experience psychological distress and disorder. Unfortunately, children in poor families are also substantially less likely to have protective factors that can counterbalance the risks, such as high-quality childcare, enrichment activities, and access to alternative learning materials (Sameroff, 2006). Moreover, children from low socioeconomic status (SES) backgrounds on average have compromised personal resources including coping skills and more poorly developed executive functioning skills, such as inhibitory control and planning (Wolff, Wadsworth, & Santiago, 2010; see “Mediators and Moderators of Poverty's Effects on Psychopathology” section). Despite these risks outlined (covered in more detail later in the chapter), there is much resilience within this population. Not all poor children succumb to school failure, experience

depression, or become involved in criminal activity. Some even thrive despite great odds. Increases in income, or emergence out of poverty arising from natural and/or governmental experiments, have been linked to declines in children's psychological problems, such as aggression (Costello, Compton, Keeler, & Angold, 2003) that persist into adulthood (Costello, Erkanli, Copeland, & Angold, 2010), while other resilient children have personal strengths, a strong adult mentor, or family/cultural strengths and resources (e.g., Werner & Smith, 2001). Psychopathology stemming from poverty is not inevitable or irreversible, but attention to the problem of poverty and its mediators and moderators is required to help eliminate this highly damaging risk factor.

Mental Health Costs of Poverty Merikangas and colleagues (2011) recently published lifetime prevalence data on psychological disorders using the National Comorbidity Study Replication adolescent supplement sample of over 10,000 adolescents aged 13 to 18. Prevalence rates were 14.3% for any mood disorder, 31.9% for any anxiety disorder, 8.7% for attention-deficit/hyperactivity disorder (ADHD), 12.6% for oppositional defiant disorder, 6.8% for conduct disorder, and 11.4% for any substance use disorder. This nationally representative sample constitutes one of the best samples to estimate the effects of SES on rates of psychopathology. In multivariate analyses accounting for family income, parental marital status, age, gender, ethnicity, and urbanicity, parental education level of less than a college degree increased the odds of all of the preceding classes of mental disorder by an overall average odds ratio of 1.6. This means that, even after accounting for a host of potential confounds, children from low-SES households were 1.6 times more likely to meet diagnostic criteria for a mental disorder than were children with college-educated parents. Unfortunately, with increasing rates of child poverty (child poverty in the United States rose from 17% to 22% between 2008 and 2012), the costs of these SES–mental health disparities are likely to continue to increase. Estimates of the economic costs of mental health problems among the poor need to take into account not only the cost of mental health care but also the costs incurred in related areas, such as psychiatric emergency room visits and inpatient stays, academic failures, crime and delinquency, and lost productivity in adulthood. Greg Duncan and colleagues (Holzer, Schanzenbach, Duncan, & Ludwig, 2008) reviewed the results of scores of rigorous research studies examining a variety of poverty-related outcomes to estimate the monetary costs to society of childhood poverty in the United States. Factoring out a conservative estimate for genetic contributions, they estimated that the experience of growing up in or near poverty, which at the time took into account approximately 17% of children (a rate that since has gone up considerably) creates a yearly opportunity cost of 1.3% of the U.S. gross domestic product (GDP), which at the time was about $170 billion per year. Opportunity costs are lost earnings resulting from reduced productivity and lower wages in adulthood as a result of growing up in poverty. Duncan and colleagues also estimated the monetary costs associated with crime and delinquency stemming from poverty, again factoring out a conservative estimate for genetic contributions. This estimate of the costs of crime, including incarceration and costs to crime victims, accounted for another 1.3% of the GDP, or another $170 billion annually. Finally, they

computed the likely economic impact of child poverty on the incidence of poor physical and mental health and their associated costs, which amounted to 1.2% of GDP and approximately $156 billion. The authors concluded that their “best estimates suggest that childhood poverty imposes costs on American society equal to nearly 4% of GDP, or about $500 billion per year” (p. 41). Thus, poverty results in enormous losses of human and economic capital every day.

Ethical Issues and Social Justice Although a thorough discussion of the causes and consequences of poverty is well beyond the scope of this chapter, it is important to place our findings within a socio-historical context. A well-documented SES–health gradient begins during childhood and persists throughout adulthood; unsurprisingly, those in poverty and with the lowest incomes have the worst health (Adler & Stewart, 2010; Braveman, Egerter & Williams, 2011). What is more surprising is that the gradient in physical health exists at all SES levels, with health deteriorating even when moving from the wealthiest SES-income level to the second wealthiest. For mental health, the same is true for the middle and bottom portions of income, but Luthar's research has shown that risk for substance use and mental health problems also show elevations in children from the most affluent homes (Luthar, 2003). The full-spectrum health gradient exists in most modern industrialized societies today. The steepness of the gradient (or the severity of health loss occurring between SES levels) varies across countries, and what determines the steepness of the gradient is the degree of income inequality in a country (Wilkinson & Pickett, 2009). In other words, the larger the gaps between the wealthiest and poorest in a society are, the steeper is the relationship between income and health. Accordingly, the United States, which has one of the largest gaps between wealthy and poor, also has one of the steepest SES-health gradients in the world. Children, who are not responsible for household incomes or economic recessions, bear a disproportionate health burden from inequality. For example, many children in the United States regularly experience food insecurity. The most recent report by the U.S. Department of Agriculture showed that 20% of households with children are characterized as food insecure, meaning that members of those households are unable to consistently access an adequate amount of nutritious food necessary for a healthy life (Coleman-Jenson Nord, Andrews and Carlson, 2011; McLaughlin et al., 2012). The effects of malnourishment on children's developing brains and bodies are well documented, and even “moderate nutritional vulnerability,” such as that found in the United States, can interfere with cognitive development (Jyoti, Frongillo, & Jones, 2005). In addition, being hungry on a regular basis interferes with the ability to concentrate and sleep and causes psychological and physiologic stress to the developing organism. Research has documented serious psychological and emotional consequences of food insecurity (Chilton, Chyatte, & Breaux, 2007; Cook & Jeng, 2009). Behavior problems, hyperactivity, aggression, social withdrawal, peer problems, and school suspension are all elevated in children who experience food insecurity, as are depression, suicidal behaviors, and anxiety (e.g., Slopen, Fitzmaurice, Williams, & Gilman, 2010). Physical health is impaired, resulting in more days sick, more days absent from school,

learning problems, and school failure, all of which can exacerbate psychological problems (Cook & Jeng, 2009). Moreover, the stress of financial strain on parents in conjunction with vigilance and anxiety about adequate food for children likely affects parental mental health in negative ways that have further repercussions for their children. Poverty also robs children of opportunities. As highlighted by the striking economic costs of childhood poverty estimated by Duncan and colleagues and further examined in Duncan, ZiolGuest, and Kalil (2010), adults who experienced poverty during their early childhoods had numerous negative work and health outcomes compared to adults whose early childhoods were not marked by poverty. On average, those who grew up poor completed 2 fewer years of schooling, worked 451 fewer hours per year, earned less than half as much income, had poorer health and higher psychological distress, were more likely to receive food stamps, and had elevated body mass indexes and overweight status. In addition, poor males were twice as likely to be arrested and poor females were six times as likely to bear a child out of wedlock compared to nonpoor counterparts. Hence, children who grow up poor are less likely to have financially rewarding careers, enjoy good health, and avoid involvement with criminal and/or social service institutions as adults.

U.S. Poverty Demographics There are approximately 72 million children under the age of 18 in the United States at this writing. Of those, 45% (32.4 million) live in homes classified as low income or near poor, and 22% (16.1 million) live in homes with incomes at or below the federal poverty threshold (FPT). Poverty rates are even higher for ethnic minority children, with poverty rates of 34% for Hispanic and 39% for African American children. The official FPT is widely recognized to be inadequate to meet the daily needs of most families, and even the federal government provides supplemental food assistance to families up to 185% of the FPT. Research has shown that many families with incomes up to 200% of the federal poverty line cannot make ends meet even with federal in-kind subsidies such as WIC (Women Infants and Children supplemental food subsidy), SNAP (Supplemental Nutrition Assistance Program, formerly known as food stamps), and the school lunch program. Taking such information into account effectively doubles the number of children facing risk from inadequate financial resources, especially for younger children and children of color. Children under the age of 6 represent 33% of the U.S. population and 48% of the low-income population (11.5 million). Rates decline slightly as children age, with rates of 45% (10.9 million) and 41% (10 million) for 6- to 11-year-olds and 12- to 17-year-olds, respectively. Rates also differ by ethnicity, with 65% of African American (6.5 million) and Hispanic (11 million), 63% of American Indian (0.4 million), 32% Asian (1 million), and 31% of white (12.1 million) children living in homes meeting lowincome criteria (Addy, Engelhardt, & Skinner, 2013). Despite programs such as Medicaid and the Children's Health Insurance Program (CHIP), 11% of children in low-income families have no health insurance.

Poverty Metrics Poverty and the related but separable constructs of economic inequality, low SES, and low

income have been measured in a variety of manners (Diemer, Mistry, Wadsworth, Lopez, & Reimers, 2013). Much scholarship in this area has focused on family-level income, social class, or SES, but increasing research is capturing poverty in a broader context of a child's life, such as in the school and in the neighborhood. We briefly review all three types of metrics. In this chapter, we use the terms poverty, low income, and low SES somewhat interchangeably because findings are generally consistent across these different ways of indexing economic hardship and/or low social position. In our reviews of the research on internalizing and externalizing and on mediators and moderators, we have tried, however, to report findings along with the specific index used in the study. Much psychological research has used either an index of social standing, such as the Hollingshead Index of Social Position, or an income-based indicator of poverty and/or wealth. Both measurement approaches have yielded important information about the size and direction of poverty's effects as well as mediators and moderators of these effects. The most basic measure is income. Duncan and colleagues (e.g., Duncan & Magnuson, 2002a) have written extensively about both the utility of income as a predictor of health and well-being and best practices for measuring income. They recommend, for example, using multiple years of income rather than a single year, because of the instability of incomes, especially among those in lower wage brackets and seasonal workers, for example (Magnuson & Duncan, 2002b). They have also developed interview questions and techniques designed to help respondents provide the best estimate of their wages (e.g., Duncan & Petersen, 2001). An alternative way of indexing family poverty is using income thresholds originally developed by the U.S. government for the purposes of determining eligibility for public assistance programs and for tracking census data. Measures of absolute poverty (e.g., the FPT) identify the amount of income minimally required to support a family (Iceland, 2003; Roosa et al., 2005). The FPT, which is designed to capture the cost of maintaining a “minimally adequate diet,” consists of a set of thresholds that vary by family size and members' ages compared to a family's pretax income (DeNavas-Walt, Proctor, & Smith, 2012). A newer supplemental poverty measure (SPM) has been developed, which takes into account the cost of clothing, shelter, and utilities in addition to food, and allows for geographic adjustments in housing costs, for example (U.S. Census Bureau, 2013). Alternatively, a child's Hollingshead Index is, for example, calculated by combining the parent(s)' educational attainment and the prestige level of their usual occupation. Such composite indices generally correlate with income, seem to be most useful in predicting family processes that mediate poverty's effects on psychopathology, and may be especially useful in samples with restricted income ranges (e.g., Wadsworth et al., 2005). Another measure of family poverty is the income-to-needs ratio (INR), which is computed by dividing total family income by the FPT for a given year and family size. INR thereby adjusts for the number of family members requiring economic resources in a household and the FPT for that year. INR cutoffs have been proposed to nominally classify the SES continuum, including families who experience extreme poverty (i.e., an INR of .50 or lower; approximately $11,500 for family of four in 2012 using the FPT; Census Bureau, 2012), low income or near poor (i.e., an INR between 1 and 2; roughly between $23,100 and $46,000 for

a family of four in 2012); and affluent (i.e., an INR greater than 4; income greater than $92,200 for a family of four in 2012) (Roosa et al., 2005). Indices of family material deprivation and hardship (Gershoff, Aber, Raver, & Lennon, 2007; Iceland, 2003; Mayer & Jencks, 1989) capture the most bare-bones basic necessities of life: food, shelter, and clothing, or lack thereof. These measures therefore tend to be most useful in samples with incomes restricted to the poor and near-poor ranges because higher-income families are unlikely to be homeless or hungry. Measures of material hardship have been useful in identifying mediators within impoverished samples, such as a parent's ability to provide educationally rich experiences and intellectually stimulating environments for their children. Material deprivation can be captured via a sum of hardships, such as housing, food, and medical care, and can be added to broader indices of cumulative risk, such as those captured by Evans and colleagues. Such cumulative risk measures expand the range of outcomes linked to material hardships to physiologic markers (Evans & English, 2002) and academic, social, and behavioral functioning (Gershoff et al., 2007). At the school level, poverty can be indexed using the proportion of the student body that qualifies for or is enrolled in the school free or reduced-cost meal program. The Healthy, Hunger-Free Kids Act of 2010 subsidizes school meals and snacks for poor and low-income children, based on FPT. Children from families with incomes at or below 130% of the FPT are eligible for free meals; those from families with incomes between 130% and 185% of the FPT are eligible for reduced-price meals. The proportion of a school that qualifies for free and reduced-cost school meals is a useful gauge of a school's aggregated poverty status; an individual's receipt of free or reduced-priced lunches can be used as an additional indicator of a child's economic status (as in Wadsworth & Compas, 2002). Finally, poverty can be assessed at a neighborhood level. For example, neighborhood disadvantage is a robust predictor of child academic failure, conduct problems, teen pregnancy, and symptoms of anxiety and depression, even in examinations spanning longer than a decade and when controlling for various measured risks and unmeasured confounds (e.g., Goodnight et al., 2012). A fairly straightforward approach to measuring neighborhood poverty, which is commonly employed in child development research, is to use the percentage of households in a given catchment area that are living below the poverty line. Neighborhoods with 30% to 40% of households with incomes at or below the FPT are generally considered “impoverished” (Leventhal & Brooks-Gunn, 2011). More complex indicators of neighborhood quality have also been calculated by capturing multiple risks within a catchment area, such as the percentage of people in a neighborhood with less than a high school education, percentage of female-headed households, violent crime rates, and percentage of unemployed adults (e.g., Santiago, Wadsworth, & Stump, 2011). Subjective assessments of the quality of a neighborhood from the perspective of current residents can also be used as indicators of neighborhood poverty, ranging from single items asking how “safe” a neighborhood is to multiple-item scales assessing social disorganization or low collective efficacy (e.g., Kohen, Leventhal, Dahinten, & McIntosh, 2008). While measures capturing multiple indicators of a neighborhood's quality tend to outperform single-item indicators, measures capturing duration of residence in addition to quality of residence tend to capture the clearest picture of

neighborhood poverty (e.g., Wodtke, Harding, & Elwert, 2011). Although neighborhood deprivation relates to children's mental health, its effects tend to be much smaller in comparison to family indices of deprivations, such as income or SES (Leventhal & BrooksGunn, 2000).

Interrelations of Poverty and Psychopathology with Other Important Life Areas In addition to psychopathology, poverty and low SES are strongly associated with adverse outcomes in other spheres, such as school readiness (Galindo & Fuller, 2010; Isaacs & Magnuson, 2011; Kingston, Huang, Calzada, Dawson-McClure, & Brotman, 2013). In their own right, academic failure, health problems, and social problems can lead to and exacerbate mental health problems, and vice versa. For example, poverty status was associated with lower cognitive skills and material hardship was associated with more social problems for children in a representative sample of 1,292 children in six rural counties in the United States (Kainz, Willoughby, Vernon-Feagans, & Burchinal, 2012). Not only does family poverty and income predict children's school readiness, but neighborhood poverty also predicted decrements in a composite measure of school readiness over time in three samples of Canadian kindergarteners (Cushon, Vu, Janzen, & Muhajarine, 2011). Similarly, neighborhood economic hardship was a significant predictor of children's lower mathematics and letter knowledge, academic outcomes, and social skills in a sample of 1,004 ethnically and geographically diverse 4-year olds, even after accounting for maternal education (an indicator of family SES) (Hanson et al., 2011). Finally, concentrated neighborhood poverty during the kindergarten years had a persistent negative effect on the reading comprehension outcomes of a sample of 2,648 urban Canadian children (Lloyd, Li, & Hertzman, 2010). Indicators of physical health, illness, and disease are also patterned by SES in children. In a sample of 8,800 children followed to 48 months of age (the Early Childhood Longitudinal Study, birth cohort), poverty increased children's odds for chronic health problems, low cognitive achievement, poor social skills, and elevated behavior problems. (Hillmeier, Lanza, Lansdale, & Oropesa, 2012). Mental health relevant biomarkers such as cortisol also show patterning by SES, as highlighted by Blair, Raver, Granger, Mills-Koonce, and Hibel (2011). They found that poor housing quality and perceived economic insufficiency were both related to children's cortisol levels in the first 4 years of life, including accumulating physiologic indicators of allostatic load. Similarly, cumulative hardship based on adequacy of food, housing, and energy predicted significantly reduced odds of health and wellness in terms of hospital visits; developmental risk; and body size, weight, and proportions in a sample of 7,141 children 4 to 36 months of age (Frank et al., 2010).

Direction of Causality Considerable debate has waged regarding the direction of the relationship between poverty and psychopathology. Social causation, or what are sometimes referred to as the social determinants of health, clearly proposes that poverty leads to psychopathology. Social

selection, however, acknowledges that individuals' own characteristics, such as their psychopathology, contributes to poverty, as a result of interference with educational and occupational attainment, for example. Consistent with selection theories, is the possibility that a third variable, such as genetics, leads to both pathology and poverty—perhaps IQ, temperament/personality, or disease-specific genes. Social selection arguments are less compelling when considering children, because children by and large do not determine their family's SES level. Here we briefly review alternative causative models. The first model, referred to as the social selection or downward drift hypothesis, posits that adults with psychological or physical health problems drift down the SES ladder due to their problems and their resulting inability to fulfill expected role obligations. The initial causes of psychopathology or poor health are assumed to be factors unrelated to SES, such as genetic liabilities (e.g., Kendler & Eaves, 1986) or unfortunate circumstances, such as an accident. Findings from several studies suggest that SES differences in the rates of a few psychological disorders, including schizophrenia (Dohrenwend et al., 1992), and possibly ADHD (Miech et al., 1999), are consistent with this hypothesis. Thus, schizophrenia, for example, may impair people's ability to secure employment and financial status commensurate with that of their parents. Little research has addressed the possibility of downward drift with regard to physical illnesses, although the few that have looked have not found evidence in support of such selection effects (e.g., Lynch, Kaplan, & Shema, 1997). The second model arises from observations that people living in poverty are more likely than those not living in poverty to engage in a variety of unhealthy behaviors—behaviors linked to various physical illnesses and diseases (e.g., Wray, Alwin, & McCammon, 2005). Rates of cigarette smoking, for example, are higher among the poor than the nonpoor, and the linkages between smoking and diseases such as cancer and heart disease are very strong (EkbergAronsson, Nilsson, Nilsson, Löfdahl, & Löfdahl, 2007). Similarly, higher rates of overweight and obesity are found among the poor, and once again obesity is linked with a variety of poor health outcomes. This is apparently an appealing explanation—poor people smoke, drink, and overeat, and that is why they are unhealthy. Unfortunately, the data do not bear out this as a complete explanation of the link between low SES and poor health by any means. In fact, studies have generally found that health risk behaviors such as these at best explain only a quarter of the association between SES and physical health and often explain significantly less than that (e.g., Lantz et al., 2001; Månsson, Råstam, Eriksson, & Israelson, 1998). In addition, as mentioned, the robust SES-health gradient also exists for mental health problems. Yet there are no compelling reasons to suspect that smoking or overeating leads to ADHD, schizophrenia, or anxiety disorders. Thus, while health risk behaviors may add something to the picture of explaining the SES-health gradient, there is much left to be explained by some other mechanism. The third model explores a related factor that could contribute to the SES-health gradient. In the United States, millions of poor people have historically not had adequate access to health insurance. Thus, many poor individuals and families have been forced to seek medical care only in emergencies, at which point many health problems are too far advanced to treat easily. While it is intuitively appealing to attribute the gradient to inadequate health insurance, once

again, this idea is not borne out in the research that has been conducted. Studies that account for access to health care have found only small percentages of the gradient explained by health care access and usage (e.g., Sapolsky, 2004). In addition, the SES-health gradient exists in virtually every country in the world, even those with universal health care programs, such as England and Sweden. Interestingly, the gradient is steepest in countries with the greatest degree of income inequality between the rich and poor, such as the United States. Finally, the SEShealth gradient exists even for diseases whose incidence is not related to access to preventive health care, such as rheumatoid arthritis and juvenile diabetes (Sapolsky, 2004). Thus, while poor access to health care is a problem and certainly does not help the situation created by the gradient, it does not appear to be the primary mechanism of the gradient. The fourth model, called social causation, posits that poor people develop psychological and physical health problems as a result of living with poverty-related adversity. Studies comparing social causation of psychological disorders with alternative models, such as social selection, generally find strong support for the social causation of multiple disorders, including depression and anxiety (e.g., Wadsworth & Achenbach, 2005). Because of the pervasiveness and chronicity of poverty, living with poverty is grueling and demoralizing—it literally wears one down mentally and physically, as evidenced by significantly higher mortality rates for those in poverty (e.g., Rehkopf & Buka, 2006 ). Research suggests that this wear and tear is not just metaphorical; chronic adversity and stress due to poverty simultaneously dysregulate the body's physiological stress response system and reduce psychological resources for coping with stress. This allostatic load eventually weakens resistance to disease-causing agents of both the physical and the psychological variety (Kahn & Pearlin, 2006; McEwen, 1998). There is considerable support for this fourth model and especially the importance of stress to the SES-health gradient (e.g., Almeida, Neupert, Banks, & Serido, 2005). Yoshikawa, Aber, and Beardslee (2012), for example, recently concluded that the evidence in support of the social causation of psychopathology supports the causal role for poverty. This evidence stands up against a variety of potential threats to validity. These authors emphasized the importance of a wide variety of mechanisms that transmit poverty's negative effects, which is consistent with the fact of poverty's long reach to almost every form of psychopathology in existence. Using a multigenerational approach to understanding causation and selection perspectives, Conger, Conger, and Martin (2010) essentially found evidence that both are in operation, but at different points in development. Research guided by their interactionist perspective, for example, has demonstrated that characteristics of the individual prior to adulthood predict adult SES and family processes, consistent with the selection proposition, and that those social and economic family characteristics predict the functioning of the next generation, consistent with social causation. Adler, Bush, and Pantell (2012) similarly reviewed the evidence relevant to social causation and selection, concluding that “there is little doubt that low SES and its accompanying social disadvantages affect health” (p. 5). They also noted that health can affect SES, and therefore the relationship between the two is dynamic, and that efforts to improve SES, for example, will likely result in health improvements, and vice versa. Thus, current thinking on the causality issue presents a more nuanced view and draws on modern genetically informative designs and advances in neuroscience to showcase the power of the

socioeconomic environment in shaping brain structures and physiologic and psychological processes related to a wide array of mental and physical health problems. We review some of the neuroscience work on SES and the brain in the section titled “Mediators and Moderators of Poverty's Effects on Psychopathology.” Current theory and research on dynamic systems models and concepts such as experiential canalization have the potential to further clarify the directionality of effects issue. For example, the experiential canalization model (Blair & Raver, 2012) proposes that as a child develops, he or she encounters a multitude of forks in the road. The pathways available to a particular child at a certain point in time are determined by both past development and current circumstances. Some circumstances promote or create a variety of pathways that lead to further growth and positive development, whereas other circumstances constrain the number of positive pathways and present more pathways with nonoptimal outcomes. This type of interactive model is not new. What is new is the idea that once a pathway is taken, the other options can become closed off, for better or worse, and subsequent forks are dependent on the previous “selected” pathway. Such models were developed in part to explain neural plasticity and the fact that some neural development will not occur for children who do not encounter the full range of pathways at a given juncture. Taking seriously the need to consider behavior at multiple levels of analysis, these models are also useful for considering how poverty may constrain positive development in some areas or amplify negative development in others. It is in this way that the expression of genes can be environmentally determined—some circumstances prevent the protein synthesis necessary to express a particular gene or the reverse (see “Future Directions” section for further discussion on genetic aspects of poverty and child psychopathology). We believe that it is in this way that poverty lays the groundwork for maladaptive pathways from developmental forks in the road, which shape a child's life pathway, neural and otherwise.

Poverty and Psychopathology That poverty places children at risk for negative life outcomes is not a new idea. The first wave of psychological scholarship on the topic began in the 1960s, coinciding with the War on Poverty. This research mostly focused on IQ and how to improve poor children's IQs and life chances. Research expanded to include socio-emotional outcomes in the 1980s and 1990s (Luthar, 1999). A special issue of Child Development on children and poverty was published in 1994, creating the first compilation of psychological scholarship covering studies of poverty's effects on child health, contextual moderators, family mediators, and interventions. In that volume, Duncan, Brooks-Gunn, and Klebanov (1994) presented results of rigorous analyses demonstrating substantial effects of family income, neighborhood composition, and family structure in the first 5 years of life on internalizing and externalizing problems at age 5. Since that special issue publication, research on the topic of child poverty has flourished, leading to several review articles in the last few decades (e.g., Aber, Bennett, Conley, & Li, 1997; Bradley & Corwyn, 2002; Brooks-Gunn & Duncan, 1997; Evans, 2004; McLoyd, 1998). Each of these reviews focuses on slightly different sets of outcomes, but the conclusions are

generally the same—poverty and income status have large effects on IQ and achievement and substantial effects on psychopathology beginning in early childhood (Bradley & Corwyn, 2002). Although genetic contributions to internalizing and externalizing pathology have gained prominence in etiologic theories in the last several decades, most research still supports that low SES, low income, and/or poverty play an important causal role (e.g., Costello et al., 2010; Odgers et al., 2012). The effects of poverty and low SES appear as both solid main effects and in interaction with inherited characteristics, such as temperament.

Internalizing Early Childhood Few studies have examined the relation between poverty and internalizing symptoms exclusively in early childhood. Rather, several studies have demonstrated a relationship between poverty and internalizing symptoms across a wide age range albeit without isolating rates of internalizing symptoms in early childhood. For example, Beidas et al. (2012) investigated whether youth living in disadvantaged neighborhoods reported higher levels of anxiety and comorbid depression as compared to youth ages 2 to 19 not living in distressed neighborhoods. They found that those living in disadvantaged neighborhoods reported higher levels of overall anxiety, social anxiety, and somatic anxiety symptoms and comorbid depressive symptoms. However, these findings are not specific to early childhood and thus do not provide a clear demonstration of the linkage between poverty and internalizing symptoms in early childhood. Of the few studies that did investigate the impact of poverty on the expression of internalizing symptoms in children in early childhood, results overwhelming indicate a positive relationship between poverty and internalizing symptoms. These findings are consistent across various poverty measures. Using measures of family level poverty (i.e., individual income) and neighborhood level poverty (i.e., percentage of households with income levels within poverty range), Duncan et al. (1994) found that youth living in persistent poverty from zero to 5 years of age had higher levels of internalizing symptoms via parent report than those who never lived in poverty. Similarly, using housing status (i.e., housed versus homeless) of families with children in early childhood, Park, Fertig, and Allison (2011) found that children who have experienced homeless episode(s) had a significantly higher percentage of internalizing problem scale scores in the clinical range than youth who were low income but housed. Strohschein (2005) found that children whose family household income at time of first evaluation (approximately 4 years of age) was higher than average had, on average, lower than expected rates of depression than those whose household income was average. Strohschein argued that these results indicate that as income increases, child depression rates decrease. This conclusion coincides with previous and more recent research illustrating the negative impact of poverty on internalizing symptoms in early childhood. Other research has indicated that experiencing persistent poverty in early childhood has shortand long-term impacts on youth, although the findings of a positive or negative impact are somewhat mixed. For example, Eamon (2000) found that both persistent and recent poverty are

positively related to internalizing symptoms, such that youth whose families recently entered poverty or have been in persistent poverty report higher levels of sadness, anxiety, and dependency than their non-poor peers. Persistent poverty in early childhood has been found to have significant impact on the emergence and continuous expression of internalizing symptoms. Najman and colleagues (e.g., Najman, Clavarino, et al., 2010; Najman, Hayatbakhsh, et al., 2010) also found that multiple experiences of family poverty across early childhood (e.g., prenatal, 6 months, 5 years) significantly increased odds of anxiety and depression during later childhood. In contrast, Strohschein (2005) found that the positive relationship between poverty and depressive symptoms attenuated over time as children matured into middle childhood. These contrasting research findings indicate that while the negative impact of poverty on internalizing symptoms in childhood is firmly established, the long-term nature of this impact is less certain. Middle Childhood Numerous studies with school-age children have found positive relations between poverty and internalizing symptoms in both cross-sectional (e.g., Reising et al., 2013; Tracy, Zimmerman, Galea, McCauley, & Vander Stoep, 2008; Wadsworth & Santiago, 2008) and longitudinal studies (e.g., Slopen et al., 2010; Spence, Najman, Bor, O'Callaghan, & Williams, 2002; Wadsworth, Rindlaub, Hurwich-Reiss, Rienks, Bianco et al., 2013). Longitudinal studies are particularly valuable due to their ability to clarify whether poverty status and patterns of poverty predict changes in children's internalizing symptoms over time. Longitudinal studies typically examine children at two points in time (e.g., Krishnakumar & Black, 2002; & Slopen et al., 2010), but several studies assessed children at three or more time points that span developmental periods including middle childhood (Ackerman, Brown, & Izard, 2004; Spence, 2002; Wadsworth et al., 2013). For example, findings from an ethnically diverse sample of children assessed every 2 years between prekindergarten and fifth grade showed that children who were poor in third and fifth grade had higher levels of internalizing symptoms, even when controlling for initial levels of symptoms (Ackerman et al., 2004). Studies that target school-age children vary according to their conceptualization and measurement of poverty which, in turn, influence results. Most commonly, poverty status is defined based on parent-reported family income and is calculated as an income to needs ratio (e.g., Brody, Stoneman, Flor, McCrary, Hastings et al., 1994). However, others have defined poverty status through receipt of welfare benefits (Takeuchi, Williams, & Adair, 1998), economic strain or financial pressure (Wadsworth et al., 2013), or a combination of povertyrelated variables, such welfare receipt, single-parent status, education, and household overcrowding (Krishnakumar & Black, 2002). The combination of poverty-related variables measured when children were 5 years of age did not predict changes in internalizing symptoms the following year (Krishnakumar & Black), suggesting that measures of poverty that are more closely related to income are better predictors of internalizing symptoms in school-age children. Other studies have found that poverty-related stress is associated with internalizing disorder symptoms and diagnoses (Wadsworth et al., 2008) and that both welfare and the inability to meet financial obligations independently contribute to internalizing symptoms

(Takeuchi et al., 1998). These findings suggest that different features of poverty make different contributions to the prediction of internalizing symptoms, and multiple features should be independently examined within the same study to further clarify these patterns. Several studies examining the relation between poverty and internalizing children in schoolage children have taken a developmental psychopathology perspective examining patterns of poverty throughout different stages in children's development (e.g., Pachter, Auinger, Palmer, & Weitzman, 2006; Slopen et al., 2010; Takeuchi et al., 1998). For example, McLeod and Shanahan (1993, 1996) found that histories of poverty in early childhood were not associated with children's internalizing symptoms at ages 4 and 5 but did predict higher trajectories of symptoms through age 10. Furthermore, these authors found that poverty history between the ages of 4 and 10 was not associated with internalizing symptoms. These findings suggest that poverty in early childhood has a latent but significant impact on the course of internalizing symptoms throughout childhood. In contrast, Spence et al., (2002) found that chronic poverty from 6 months of age through 14 years of age predicted parent- and child-reported internalizing symptoms for girls only. These two studies differed in their definitions of poverty and their age ranges, which might in part explain their discrepant findings. The strength of the relation between poverty and internalizing symptoms also depends on how internalizing symptoms are measured. The majority of studies measuring this relation in school-age children use parent-reported questionnaires of symptoms (e.g., Pachter et al., 2006; Slopen et al., 2010). Other studies have created composite variables using parent and child reports of symptoms (e.g., Reising et al., 2013; Spence et al., 2002), teacher-reported symptoms (Ackerman et al., 2004), or mother and father reports of depressive symptoms (Brody et al., 1994). The few studies that have used diagnostic interviews have shown poverty to be related to clinical disorders and symptom severity (Costello et al., 2003; McLaughlin et al., 2012; Tracy et al., 2008). Future research in children of this developmental time frame should focus on utilizing children's self-reports, which can be reliably obtained in children as young as 8 years old (e.g., Tilghman-Osborne, Cole, & Felton, 2012; Twenge & NolenHoeksema, 2002). Adolescence Several studies have demonstrated mixed findings for a significant relationship between neighborhood poverty and internalizing symptoms in adolescents. In their comprehensive review of the relationship of neighborhood poverty on adolescent development, McBride Murry, Berkel, Gaylord-Harden, Copeland-Linder, and Nation (2011) found some support for direct effects of neighborhood poverty on internalizing symptoms in youth. For example, youth from low-SES families reported higher levels of depressive symptoms than youth not living within low-SES families (Wickrama & Bryant, 2003). In contrast, some research findings did not demonstrate a significant relationship between poverty and internalizing symptoms in youth. For example, in their investigation of mediating factors in the relationship between neighborhood disadvantage (as measured by indicators of neighborhood poverty) and adolescent internalizing problems, Deng and colleagues (2006) did not find that residing in a disadvantaged neighborhood was associated with reported youth internalizing symptoms.

These contrasting findings suggest that the link between poverty and internalizing symptoms may be different based on level of poverty experienced by youth (neighborhood versus individual) and/or in combination with family poverty levels. Similar to research in early childhood, length of poverty experience has a significant impact on the presentation of internalizing symptoms in youth. For example, Slopen and colleagues (2010) investigated the effects of persistent poverty on children at two time points (baseline 4 to 14 years of age and follow-up 5 to 16 years of age) and found that, at follow-up, those youth whose family reported continued food insecurity were 1.5 times more likely to experience internalizing symptoms. Other dynamic relationships between poverty and internalizing symptoms in youth have been found. Leventhal and Brooks-Gunn (2011) found that in highpoverty neighborhoods decreasing poverty was, surprisingly, associated with an increase in male internalizing problems, while in moderate-poverty neighborhoods, increasing poverty was associated with an increase in male internalizing problems. Similar effects were not found for female youth.

Externalizing Early Childhood Much research on psychopathology in young children has focused on behavioral and emotional dysregulation rather than psychiatric diagnoses. In numerous studies, toddler and preschooler behavior problems occur at significantly higher rates in samples of low-income children (e.g., Barker, Copeland, Maughan, Jaffee, & Uher, 2012; Dearing, McCartney, & Taylor, 2006; Yoshikawa et al., 2012). The association exists in studies using umbrella terms such as conduct problems, behavior problems, and undercontrolled behavior as well as studies examining more specific behaviors, such as aggression, noncompliance, hyperactivity, and oppositionality. Moilanen, Shaw, Dishion, Gardner, and Wilson (2010), for example, examined predictors of the growth of inhibitory control (protective against externalizing problems) from 2 to 4 years of age. Extreme family poverty predicted significantly slower growth in inhibitory control in this sample of 731 youngsters. Barker and colleagues (2012) found, for example, that contextual risk (e.g., poverty) during pregnancy predicted children's subsequent age 0 to 2 behavioral dysregulation (a composite of hyperactivity, conduct problems and emotional difficulties); contextual risks during the child's first 2 years of life predicted subsequent age 4 behavior dysregulation. Henninger and Luze (2013) estimated latent growth curve models of predictors of externalizing symptoms among early Head Start enrollees (ages 0–3). A primary pathway to externalizing in this population was via poverty, and the authors identified parental stress and depression as mediators of this pathway. Gershoff and colleagues (2007) examined the ability of both income and material deprivation to differentially predict social-emotional functioning in the Early Childhood Longitudinal Study sample of 21,255 kindergarteners. They found that the combination of both family income and material hardship was the strongest predictor of child competencies (a latent factor consisting of internalizing, externalizing, and social skills). Some research has focused on the presence and stability of clinically elevated externalizing

symptoms in toddlers and preschool-age children. Hill, Degnan, Calkins, and Keane (2007), for example, found low SES at age 2 to be a significant predictor of chronically elevated clinical levels of externalizing symptoms at ages 2, 4, and 5. This type of chronic course of clinically elevated externalizing symptoms, referred to as the early starter pathway, tends to occur much more frequently in low-SES families. The early starter pathway to disruptive behavior disorders is a particularly troubling pathway that portends lifelong disorder (Campbell, 1995; Hinshaw, Lahey, & Hart, 1993). Toddlers in general engage in high levels of aggressive, antisocial behaviors that tend to desist over time as they acquire better communication abilities and ways of handling conflicts and negative emotions (Coie, Dodge & Damon, 1998). There is a subgroup of children, however, that does not desist over time and rather stays elevated into middle childhood and beyond, ultimately meeting diagnostic criteria for disruptive behavior disorders such as conduct disorder. Poverty and low SES are strong predictors of membership in this early starter pathway. Keller, Spieker, and Gilchrist (2005), for example, found that an index of ecological risk, which included family economic distress and parenting stress, predicted membership in an early starter pathway of increasing conduct problems over time at twice the rate expected by chance, specifically for children with insecure attachments. Studies have increasingly used rigorous quasi-experimental and genetically informed designs to isolate the causal contribution of poverty and SES to the development of externalizing problems. D'Onofrio et al. (2009), for example, incorporated comparisons of target children with genetically related and unrelated children to rule out a variety of unmeasured environmental and genetic confounds in their estimates of the causal effects of income on conduct problems in children as young as age 4. These authors found strong evidence in support of the inference that family income is inversely and causally related to conduct problems, especially for boys. Middle Childhood Rates of diagnosed externalizing disorders rise during the school years, in part due to children's entrance to school and the increasing compliance demands. School also often creates a context in which underlying liabilities, such as ADHD, create additional problems with peers, teachers, and parents that can contribute to further behavior problems. Lansford et al. (2006) found consistent effects of low SES on the development of externalizing trajectories over time, from kindergarten to the eighth grade, in their community sample of 585 children with and without physical abuse histories. Similarly, Slopen and colleagues (2010) found robust effects of food insecurity (highly correlated with poverty) on externalizing problems of a sample of 2,810 children aged 4 to 14 followed for 1 year. Church, Jaggers, and Taylor (2012) used data from the Fragile Families Project to test a variety of predictors of child behavior problems in this at-risk sample of more than 5,000 children. Family poverty level was a robust predictor of behavior problems, even while simultaneously accounting for neighborhood SES, gender, parenting, and parental stress. Parental stress partially mediated this pathway. Using data on 3,259 families with a total of 5,808 children from the National Longitudinal

Survey of Youth, Hao and Matsueda (2006) examined the effects of concurrent poverty as well as poverty during early childhood on externalizing behavior problems in middle childhood. They found strong evidence of negative effects of early childhood poverty on subsequent externalizing symptoms, an effect that was not mediated by current parenting behaviors. Odgers and colleagues (2012) examined the differential patterning of antisocial behavior by SES level over time in a large, nationally representative British birth cohort of 2,232 children followed from age 5 to 12. They found that rates of antisocial behavior did indeed show a patterning by SES, such that children in the “deprived” SES group had significantly higher rates of antisocial behavior than did middle-SES children, who in turn had significantly higher rates than highSES children. These disparities persisted from age 5 to age 12 and widened over time, such that the mean difference in levels of antisocial behavior between the deprived and more affluent SES groups increased from d = 0.38 at age 5 to 0.51 at age 12. Odgers and colleagues also found that supportive parenting mediated the relationship between both neighborhood and family-level SES and child antisocial behavior. Finally, Cote, Vaillancourt, LeBlanc, Nagin, and Tremblay (2006) followed 10 cohorts of children for 6 years and examined trajectories of physical aggression over the period from 2 to 11 years of age. They identified three subgroups of children, two of which exemplified a typical developmental pattern of infrequent and declining physical aggression over time. The third group of children contained mostly boys from disadvantaged families and these boys exhibited an atypical developmental pattern consisting of more frequent and stable use of physical aggression. Adolescence The effects of adverse environmental circumstances on the development of adolescent disruptive behavior problems are undeniable (Moffitt, 2005; Paschall & Hubbard, 1998; Thornberry, Freeman-Gallant, Lizotte, Krohn, & Smith, 2003; Wiesner & Windle, 2004). Several studies have followed children from infancy or toddlerhood into adolescence. Such studies provide strong support for causal linkages between SES and externalizing pathology. Schonberg and Shaw (2007) examined early childhood predictors of trajectories of conduct problems into early adolescence, finding that children who experienced stable poverty in early childhood were at significantly elevated risk for being classified in a chronic conduct problems group as compared to children who did not experience poverty or who experienced only temporary poverty. Fanti and Heinrich (2010) found high levels of sociodemographic risk (low maternal education, low income) predicted high and continuing levels of conduct problems in the National Institute of Child Health and Human Development (NICHD) Study of Early Childcare sample of 1,364 children followed from age 2 to age 12. Dearing and colleagues used this same data set to examine within-child associations between family income and externalizing symptoms. They also found large main effects of chronic poverty and smaller effects of income on externalizing symptoms, but, importantly, chronic poverty and income interacted such that chronic poverty amplified the effects of low income on externalizing (Dearing et al., 2006). Finally, Stroschein (2005) used the National Longitudinal Study of Youth, Child Supplement, sample of 7,143 children age 4 to 14 to examine the dynamic effects

of income and income changes over time on children's antisocial behavior. Her analyses showed that low income at entry into the study was associated with both initially high levels of antisocial behavior and with increasing levels and rates of change in antisocial behavior over time. Additionally, increases in income were associated with decreasing antisocial behavior and vice versa. Bøe, Øverland, Lundervold, and Hysing (2012) examined the effects of several different indicators of SES in predicting externalizing disorders in a large sample of 11- to 13-year-olds (N = 5,781), and found that each of their SES indicators consistently predicted externalizing disorders. Analyses of data from the Mater-University Study of Pregnancy on 3,103 children followed from birth to age 21 revealed very strong effects of poverty on persistent aggressive and delinquent behavior (Najman, Clavarino, McGee, Bor, et al., 2010). Family poverty experienced at the 14-year follow-up, for example, predicted persistent aggression, delinquent behavior, smoking, and alcohol use at the 21-year follow-up. Additionally, the largest effects were found for children with recurrent experiences of family poverty. Children who were classified at three or four of the five assessments as living in poverty were more than twice as likely to be aggressive and delinquent at both 14 and 21 years. A compelling new study by Boyce and colleagues recently examined rates of physical and mental health problems using several different types of indicators of SES (Nuru-Jeter, Sarsour, Jutte, & Boyce, 2010). By every metric examined, externalizing problems were inversely associated with SES. A unique feature and finding of this study is that associations involving categorical representations of SES were much more pervasive and stronger in magnitude than were measures capturing monotonic gradations in SES. The authors suggested that the standard use of income and maternal education in research probing the SES-health phenomenon may be underestimating the strengths of the SES-health association by using these “less impactful” measures. The effects of low SES and poverty have also been examined among particular high-risk groups of children. In their study of children of depressed parents, for example, Reising and colleagues (2013) found that economic disadvantage predicted externalizing symptoms, even among this at-risk group. Disrupted parenting was a significant moderator of this effect. In Kjellstrand and Eddy's (2011) study, children of incarcerated parents had consistently higher levels of problem behavior between fifth and tenth grade than children without an incarcerated parent. The strength of this association increased over time. A growing number of studies have documented that poverty or disadvantage at the neighborhood level is also strongly associated with the development of aggression, violence, and delinquency in adolescents (e.g., McBride-Murry et al., 2011). Exposure to deviant peers and lack of collective efficacy have been offered as potential mechanisms of this effect (e.g., Haynie, Silver, & Teasdale, 2006), and gender sometimes is found to moderate these effects (e.g., Kroneman, Loeber, & Hipwell, 2004). Harrington (2003) used behavioral genetic methods to estimate genetic and environmental contributors to aggression and delinquency in the Add Health project sample of 2,434 sibling pairs of children. First, they found that aggression and delinquency were significantly higher in disadvantaged neighborhoods than in advantaged neighborhoods. In addition, they found significant shared environmental effects on aggression and delinquency, but only in the disadvantaged neighborhood context, “providing

general support for the idea that disadvantaged neighborhood contexts directly affect developmental influences on adolescent aggression” (p. 233). Karriker-Jaffe, Foshee, Ennett, and Suchindran (2009) found that both neighborhood disadvantage and social disorganization predicted adolescent aggression trajectories over 2.5 years for boys, whereas only neighborhood disadvantage predicted aggression for girls.

Mediators and Moderators of Poverty's Effects on Psychopathology In this section of the chapter, we describe other constructs, such as stress and coping, that likely play a role in the relations between childhood deprivation and mental health. Much of the work has focused on underlying pathways or mediators that help account for the robust SES–childhood pathology link. Three predominant examples of mediators of the SES– psychopathology link in children are disrupted parenting, elevated stress, and alterations in executive functioning. Other constructs factor in as moderators or variables that alter children's vulnerability to deprivation. Gender, ethnicity, and coping strategies are the three most studied moderators of the relations between SES and children's mental health. There are even a few rare instances of moderated mediation whereby the underlying pathway linking SES to child psychopathology is moderated by some other variable. For example, we show that a major pathway linking SES to psychopathology is parenting problems created and exacerbated by chronic poverty-related stress. The dynamics of this pathway can be altered by social support. In the “Future Directions” portion of this chapter, we outline some other mediating constructs warranting attention in investigations of SES and child psychopathology.

SES and Parenting Low-SES parents tend to be harsher and less responsive to their children—theoretical and empirical models have suggested that these parenting problems do not necessarily reflect underlying parenting deficits (though they may in some cases) but rather that the stress and chaos of life in poverty interferes with the ability to deliver warm, responsive parenting (Conger, Conger, & Martin, 2010). These features of parenting are strongly and consistently related to concurrent and future mental health problems in children (Bradley, 2003; Repetti, Taylor, & Seeman, 2002). A large number of studies support that stress, especially economic strain and family conflict, accounts for the association between SES and parenting, as illustrated in Figure 4.1. Beginning with early work examining the role of financial difficulties during the Great Depression and child development (Elder, 1974), considerable research has uncovered support for this model. Rand Conger and colleagues, including Elder (Conger & Elder, 1994), have extended this model, suggesting that financial strains related to insufficient income for families leads to emotional distress in parents as well as interparental discord, both of which lead to more negative parenting (Conger & Donnellan, 2007).

Figure 4.1 Conceptual model of SES and parenting. Evidence supporting the model depicted in Figure 4.1 is impressive, including both crosssectional and longitudinal data that cut across ethnic samples within the United States and across different countries (for reviews, see Bradley, 2003; Conger & Donnellan, 2007; Grant et al., 2003; Hoff, Laursen & Tardif, 2002; Magnuson & Duncan, 2002b). There is also a small amount of experimental research showing that supplements to family income can lead to increased parental warmth and responsiveness (Costello et al., 2003; Gennetian & Miller, 2002; Morris, Duncan, & Clark-Kauffman, 2005), presumably via reductions in povertyrelated stress. In a meta-analysis of nearly 50 studies on family SES, parenting, and youth development, Grant and colleagues (2003) found an overall effect size of poverty on negative parenting of d = .48 (cross-sectional evidence) and d = .55 (longitudinal evidence). Analyses indicated that poverty had both direct and indirect effects on both internalizing and externalizing symptoms in children and youth. Grant et al.'s conclusion that a partial mediation model is the best fit to the data is in line with findings reviewed later that low-SES environments have direct effects on children. Furthermore, it is likely that the effects of a low-SES childhood are due to the accumulation of exposure to multiple risk factors. In terms of Figure 4.1, there is good evidence of additional, parallel mediating factors, which operate similarly to negative parenting, that are also capable of accounting for some of the ill effects of poverty on children's mental health. A critical risk factor for negative mental health outcomes in children is maternal depression (Downey & Coyne, 1990; Green et al., 2010; see also Chapter on Depression/Mood Disorders. in this handbook). Low-income adults, especially women, have higher rates of depression (Belle, 1990; Green et al., 2010). Given abundant documentation that depressed maternal affect and poverty are both linked to diminished parental responsiveness, more work needs to be done on the potential critical role of parental mental health and particularly maternal depression in the relations between childhood poverty and children's development (Shonkoff & Garner, 2012). Across several studies, parents experiencing greater financial strain suffer greater emotional distress, including symptoms of depression, with a median path coefficient of .42 (Conger & Donnellan, 2007). In an interesting set of results, Harnish and colleagues (1995) showed that the inverse link between SES and externalizing symptoms in first graders was mediated by maternal responsiveness and warmth. Much of the negative covariation between SES and parenting was caused by elevated maternal depression. Among a sample of younger children (28–50 months), the adverse impacts of SES on behavioral adjustment were accentuated for those whose mothers were more depressed (Petterson & Albers, 2001).

This model of SES, parenting, and children's mental health is also robust with respect to ethnicity (Conger & Donnellan, 2007; Eccles & Gootman, 2002) and gender (Grant et al., 2007), although some research suggests enhanced benefits of warm parenting for ethnic minority children. Bradley and colleagues (2001), for example, showed that the link between poverty and less nurturant parenting was stronger in African American families vis-à-vis white or Hispanic families in a large, nationally representative data set. Of additional interest, a Canadian national cohort study indicates that children (ages 4–11) in recent immigrant families may experience some protection, particularly for externalization symptoms, because of more positive parenting practices (Georgiades, Boyle, & Duku, 2007). There also does not appear to be any clear pattern for ethnic variation in the direct effects of SES or income levels on psychopathology in children. In a 4-year longitudinal study of children in grades 2 through 4, Bolger, Patterson, Thompson, and Kupersmidt (1995) at the onset found that African American children manifested more internalizing symptoms as a function of persistent poverty whereas White children showed more externalizing symptoms. Costello, Keeler, and Arnold (2001) showed a stronger link between poverty and psychopathology among White relative to African American school children. An index of cumulative risk that incorporated family and neighborhood poverty along with several other risk factors that co-occur with poverty was also more strongly associated with internalizing and externalizing symptoms among White relative to African American children (Gerard & Buehler, 2004). Among young adolescents, however, Nuru-Jeter and colleagues (2010) found the opposite pattern for externalizing symptoms with stronger income effects on African American youth. Furthermore, some studies suggest that chronic poverty may have more pronounced effects among boys than girls because of harsher parenting (Bolger et al., 1995; Grant et al., 2007). Elder, vanNguyen, and Caspi (1985) also found the impacts of economic hardship were conditioned by the gender composition of the parent-child dyad. But overall, the degree of fit for the model depicted in Figure 4.1 across gender is high. Turning to gender as a moderator of the direct effects of SES on psychopathology, again the data generally indicate weak or contradictory findings. In their longitudinal study of poverty duration and psychopathology in a sample of grade-school children, Bolger and colleagues (1995) found no effects of poverty by gender for internalizing symptoms, but boys were more vulnerable for externalizing symptoms. Henninger and Luze (2013) found the opposite effect among children with girls showing greater externalizing symptoms in relation to time spent in poverty. In a sample of Australian youth followed from birth through adolescence, girls, but not boys, with early childhood exposure to poverty manifested more internalizing symptoms (Spence et al., 2002). In a longitudinal study spanning age 9 to 17, Evans and colleagues (Evans & English, 2002; Evans, Gonnella, Marcynyszyn, Gentile, & Salpekar, 2005) found no evidence of gender differences across a wide range of physiological stress, self- and parental ratings of children's mental health, and in behavioral probes of psychological health. Roosa et al. (2005) as well as Mrug and Windle (2009) found no gender differences in children's internalizing or externalizing responses to neighborhood poverty. Similarly, Nuru-Jeter and colleagues (2010) found no gender differences in young adolescents' externalizing responses to family income.

Summarizing, the studies on SES, parenting, and psychopathology indicate clear and consistent support for the pathway depicted in Figure 4.1. This robust pathway holds across gender and ethnicity and has been replicated several times internationally. The direct effect of SES or income on psychopathology also tends to be robust with no clear pattern emerging for variable vulnerability as a function of either ethnicity or gender. This is not to argue that these factors are unimportant in thinking about poverty and children and youth's mental health. Our point is simply that the SES → Problematic Parenting → Mental Health model is strong and well documented. At present, we do not have clear-cut indications of whether this pathway varies by gender or ethnicity. The preponderance of current data suggests that the effects of poverty or low SES early in life on children and later on adolescents appear to be similar across race and gender. A few studies suggest that high levels of social support may buffer the ill effects of poverty on parenting (Conger & Elder, 1994; Hashima & Amato, 1994; McLoyd, 1997). Greater religiosity may have similar protective effects (Brody, Stoneman & Flor, 1996; Grant et al., 2000). Although social support and religiosity may buffer some of the ill effects of poverty on parenting, chaos can have the opposite effect. Chaos is a composite index reflecting noise, crowding, interruptions, routines, and structure in the household. Chaotic household conditions accentuated the effects of negative parenting on preschool and elementary school children's problem behaviors (Coldwell, Pike, & Dunn, 2006). The negative impact of chaos on the executive functioning of preschool mothers was exacerbated by multiple risk factors associated with poverty (Deater-Deckard, Chen, Wang, & Bell, 2012). Thus, it is also worth thinking about not only the direct effects of low SES conditions on parents which, in turn, can alter critical parenting behaviors such as warmth and responsiveness, but also factors that might undermine parental resources. Some of the conditions that covary with poverty and deprivation, such as chaotic living situations, lower levels of social support, or food insecurity, likely also erode parental coping resources, which in turn could alter important parent-child interactions as well as the overall social climate of the family.

SES and Stress In addition to the work reviewed earlier on SES, parenting, and children's psychopathology, multiple methodological probes indicate that children themselves experience poverty as a stressful condition. Children in low-SES households report high levels of poverty-related stress, including such things as financial pressure, family conflict, family disruption and change, discrimination, and violence (Wadsworth & Berger, 2006; Wadsworth & Compas, 2002; Wadsworth &t Achenbach, 2005; Wadsworth et al., 2008; Santiago et al., 2011; Wolff, Santiago, & Wadsworth, 2009). Many studies revealed that among low-SES children, those experiencing greater poverty-related stress also had higher symptoms of internalizing and externalizing. Moreover, Santiago and colleagues (2011) found that poverty-related stress within a low-SES sample also interacted with prior mental health symptoms, aggravating their severity over time. Wolff and colleagues (2009) also showed in a prospective, longitudinal design that the impacts of poverty-related stress on adolescent mental health symptoms were more severe for those individuals with greater reactivity to stress.

The descriptive data on high levels of poverty-related stress among children and youth from low-SES households matches comparative data contrasting low-SES children relative to middle-SES children. In comparison to their middle-SES peers, low-SES children and youth are exposed to substantially more psychosocial stressors, such as family conflict and turmoil, changes in family composition, violence, parental harshness and low responsiveness (Attar, Guerra, & Tolan, 1994; Bendersky & Lewis, 1994; Brooks-Gunn, Klebanov, & Liaw, 1995; Brown, Cowen, Hightower, & Lotyczewski, 1986; DuBois, Felner, Meares, & Krier, 1994; Dubow, Tisak, Casey, Hryshko, & Reid, 1991; McLoyd, 1998; Pungello, Kupersmidt, Burchinal, & Patterson, 1996) as well as to worse physical conditions capable of producing stress, such as toxins, noise, crowding, and substandard housing (Evans, 2004). One aspect of the settings of low-SES children and youth that appears to be especially problematic is elevated exposure to multiple stressors or cumulative risk factors. A large literature documents that exposure to cumulative risk factors relative to singular ones, leads to worse developmental outcomes, including psychopathology, physiological stress, and poorer cognitive development (Evans, Li, & Whipple, 2013; Obradovic, Shaffer, & Masten, 2012; Sameroff, 2006). Children from low-income households from birth through late adolescence are much more likely to be exposed to a high number of cumulative risk factors (Burchinal, Roberts, Hooper, & Zeisel, 2000; Deater-Deckard, Dodge, Bates, & Pettit, 1998; Evans & English, 2002; Evans & Kim, 2007, 2010; Federman et al., 1996; Felner et al., 1995; Greenberg, Lengua, Coie, & Pinderhughes, 1999; Liaw & Brooks-Gunn, 1994; Rutter, 1979). They are also more likely to be exposed to chaotic living conditions indicative of a low degree of structure, stability, routine, and predictability in daily life (Evans, Eckenrode, & Marcynyszyn, 2010). Considerable work has linked high levels of family chaos with deficits in socioemotional development in children and youth (Fiese, 2006; Fiese & Winter, 2010). An important, emerging area of research on SES and health focuses on physiological indices of chronic stress as candidate mechanisms to account for well-documented SES gradients in health in both children and adults (Adler & Stewart, 2010; Braveman et al., 2011). This research is too voluminous to describe in detail so we summarize major findings from overviews of this work. Prior to adolescence, lower SES is associated with higher blood pressure (Chen, Matthews, & Boyce, 2002; Evans et al., 2012). A number of studies also reveal that childhood SES is inversely related to stress hormones in children, including cortisol, epinephrine, and norepinephrine; metabolic impairments, such as inefficient glucose metabolism, higher levels of cholesterol and low-density lipids; as well as elevated inflammatory responses indicative of disturbed immune function (Evans et al., 2012). Allostatic load, a marker of dysregulation across multiple response systems (i.e., hypothalamic-pituitary-adrenal axis, sympathetic adrenal medullary axis, metabolic function, inflammation, and immune function) is an exciting direction because of encouraging evidence that this marker of wear and tear on the body is a better predictor of physical and psychological morbidity than singular stress markers (Ganzel, Morris & Wethington, 2010; Juster et al., 2011; McEwen, 1998). Children from lower-SES backgrounds have higher levels of allostatic load (Evans et al., 2012). Summarizing, there is abundant evidence that household SES is inversely related to both

psychosocial and physical stressor exposures. One feature of this inequality that may be particularly important for children's mental health is exposure to the accumulation of multiple stressors or risk factors. Low-SES children and youth are much more likely to encounter a plethora of risk factors, both psychosocial (e.g., family turmoil, negative parenting) and physical (e.g., substandard housing, toxins), in comparison to their more advantaged peers. Finally, low-SES children manifest a consistent pattern of elevated, chronic physiological stress relative to middle-SES children. Although there is emerging work showing that chronic stress accompanying a low-SES childhood may help explain the etiology of physical morbidity later on in life (Evans et al., 2012; Evans & Kim, 2010), very little research has directly examined SES → Stress → Mental Health pathways. In a large, nationally representative sample of American kindergarten children, negative associations between household income and both cognitive skills and socioemotional competency were partially mediated by a composite index of exposure to material hardship (Gershoff et al., 2007). (Material hardship incorporated measures of food insecurity, residential instability, inadequate medical care, and duration of financial difficulties since the child's birth.) Furthermore, the pattern of findings was similar across White, African American, and Hispanic kindergartners (Raver, Gershoff, & Aber, 2007). Higher levels of cumulative risk exposure mediated much of the relation between poverty in elementary-age school children and both elevated physiological stress and psychological distress (Evans & English, 2002). Cumulative risk consisted of an index of substandard housing, residential noise, residential crowding, family turmoil, child separation from family, and exposure to violence. Chaotic living conditions among middle school children partially mediated the link between childhood poverty and elevated symptoms of internalizing and externalizing (Evans et al., 2005). Felner and colleagues (1995) showed that the inverse relation between SES and multiple measures of internalizing symptoms among middle school children was mediated by stressful events, family supportive climate, and school belonging. Similarly, eighth-grade students from lower-SES families have higher rates of substance abuse (Wills, McNamara, & Vaccaro, 1995). This finding was mediated, in part, by a composite index of multiple risk exposures (e.g., low parental support, high negative life events, deficient youth competency, and contact with peer substance abusers). In a large sample of early adolescents, AmoneP'Olak and associates (2009) noted that the inverse relations between SES and both internalizing and externalizing symptoms were mediated by exposure to negative life events outside the personal control of adolescents (e.g., housing and neighborhood problems, serious illness/injury to family members). Sameroff, Bartko, Baldwin, Baldwin, and Seifer (1998) formed a composite index of adolescent adjustment incorporating internalizing, externalizing, extracurricular/community involvement, and academic performance that was negatively correlated to family income. This significant association, however, was reduced to near zero when exposure to cumulative risk factors was included in the model. A large number of cumulative risks across the domains of maternal competence, parenting, family structure, family social networks, peer deviance exposure, and neighborhood social capital were incorporated into their model of cumulative risk exposure. African American adolescents (ninth and eleventh grade) from poorer families experienced higher levels of depression (Hammack, Robinson, et al., 2004). For girls but not boys, these relations were mediated by

higher levels of family stress (e.g., unmarried family member pregnancy, close family relative died). Finally, we also know that cumulative risk exposure appears to be an important factor underlying the robust relations between childhood deprivation and many of the indicators of chronic physiological stress already outlined (e.g., elevated blood pressure, stress hormones, allostatic load) (Evans & Kim, 2010, 2012). One study that found a different pattern of results among SES, stress, and psychological health among children was conducted by DuBois and colleagues (1994). They found a synergistic interaction between the degree of disadvantage among adolescents and exposure to daily hassles on indices of both internalizing and externalizing symptoms in a prospective, longitudinal design over the academic year among seventh and eighth graders. One possible explanation for this pattern of moderation rather than mediation indicated by the majority of studies mentioned could be the index of daily hassles used by DuBois and colleagues. They summed the ratings of hassle severity to form a composite score. This method contrasts with other studies that relied on risk exposure metrics.

SES, Executive Functioning, and Coping Given evidence that SES is associated with levels of stress in children, there are at least three reasons also to explore executive functioning and coping as mechanisms that may contribute to the relations between SES and psychopathology. First, considerable work shows that chronic stress, in both animal and human models, is capable of damaging executive functioning (EF) (Blair, 2010; Lupien, McEwen, Gunnar, & Heim, 2009; Repetti et al., 2002). EF is a complex construct that typically encompasses self-regulation, attentional control, working memory, behavioral inhibition, delay of gratification, and planning. EF is believed to lay the early foundation that undergirds subsequent development and maintenance of coping strategies invoked when stressors are encountered, and it increasingly plays an important role in theory and research regarding the development of various types of psychopathology (Hofer, Eisenberg, & Reiser, 2010; Rinsky & Hinshaw, 2011). In the next study descriptions, we provide thumbnail sketches of EF measures in order to convey a better sense of what processes actually are involved. The second reason why EF and coping are likely related to childhood SES is because many of the risks encountered early in life among low-SES children are chaotic and largely uncontrollable, both critical conditions that can undermine the development of coping skills (Evans & Kim, 2013; Lengua, 2012; Raver, 2004; Repetti et al., 2002; Wolff et al., 2011). Third, a large body of literature in the stress field indicates that coping strategies can moderate experiences of stress in children and adolescents (Aldwin, 2007). Thus, we would expect that how children cope with poverty and some of its accompanying risk factors has implications for mental health outcomes. A standard approach to the assessment of EF is to use neurocognitive testing batteries with well-characterized brain correlates in the prefrontal cortex. In a program of research on childhood SES and neurocognitive development, Noble, Farah, and their colleagues uncovered consistent evidence of EF deficiencies among low- compared to middle-SES children ranging from 5 to 13 years of age (Farah et al., 2006; Noble et al., 2005; 2007). EF was assessed by standard neurocognitive batteries including measures of inhibitory control such as reaction time on the number Stroop task (say the number shown [“2”] or say the quantity of numbers

shown [e.g., four “2s” are shown in a square, correct answer “4”]), and the flanker task, which measures reaction time to indicate what direction a stimulus (e.g., arrow) is pointing in when surrounded by adjacent stimuli pointing in the same or in opposite directions. They also found deficits in working memory. Utilizing an EF composite consisting of (1) working memory, (2) the classical Stroop task (say the color of ink a color word is written in), and (3) an index of planning, Hughes, Ensor, Wilson, and Graham (2010) found that family income was negatively associated with EF among 3.5-year-olds, but it did not influence their rate of improvement from toddlerhood to 6 years of age. Mezzacappa (2004) also found that 4- to 7-year-olds performed worse on the flanker task if they were from lower-SES families. Three- to 5-yearolds from lower-SES backgrounds manifested worse EF on two Stroop-like measures adapted for preschool children (Li-Grining, 2007). Raver, Blair, and colleagues (2013), utilizing a composite EF battery (working memory, attention control, inhibitory control) at 48 months of age, showed that the number of times families had incomes at or below the poverty line in the 3 previous years predicted EF among 4-year-olds. They also found that a parallel index of perceived financial strain had similar prospective association with EF, but in the latter case, the association between financial strain and EF was moderated by an index of infant emotional reactivity. Only infants high in emotional reactivity showed the expected negative association of early experiences of deprivation and EF deficits. Five- to 14-year-old children from Colombia and Mexico performed a standard index of attentional control in which they had to classify visual stimuli according to varying dimensions (e.g., shape, color, number). Children from lower-SES households had more difficulty with the task (Ardila, Rosselli, Matute, & Guajardo, 2005). Similarly, Lipina, Martelli, Vuelta, and Colombo (2005) found that 6- to 14-month-old infants from lower-SES families were less skilled in an attentional control task in which they had to search for a hidden object either where it was originally placed or after it had been placed in a new location. This task pulls for both working memory and attentional control. Five- to 6-year-olds from low-income families had greater difficulty resisting an interesting visual distractor while sustaining performance on a visual attention task (Howse, Lange, Farran, & Boyles, 2003). Recently a team of investigators examined how EF developed throughout preschool and kindergarten in relation to income (Wanless, McClelland, Tominey, & Acock, 2011). At the initial evaluation, lowincome 4- and 5-year-olds had more difficulty in a response inhibition task (Simon says), but, interestingly, these same children gained more inhibitory skills throughout the second half of their prekindergarten year on through kindergarten. Low-income children whose first language was not English, however, did not manifest this more rapid gain in inhibitory skills. It is also worth noting that despite the more rapid gains overall among the low-income preschoolers and kindergartners, their EF skills did not catch up to those of their more affluent peers by the end of kindergarten. Utilizing maternal ratings of inhibitory control (e.g., “Can easily stop an activity when s/he is told no.”), the EF skills of 2- to 3-year-olds were unrelated to family poverty, but the maturation of EF skills over a 2-year period was slower for poor than nonpoor toddlers (Moilanen et al., 2009). A few studies have looked directly at the possible role of stressor exposure in the linkage between SES and EF. In a series of studies, Evans and colleagues have shown that poverty is

related both cross-sectionally and longitudinally with deficits in the ability to delay gratification in toddlers (Evans & Rosenbaum, 2008) and in 9-year-olds (Evans & English, 2002) as well as teacher ratings of elementary school children's self-regulatory abilities in the classroom (Evans et al., 2005). The latter two studies also found that the poverty–delay of gratification link was mediated in part by cumulative risk exposure and chaotic living conditions, respectively. On a different measure among a sample of low-income preschool children, Li-Grining (2007), however, found no relation between SES and of delayed gratification. Evans and Schamberg (2009) showed that the prospective relation between early childhood poverty and working memory in young adults was mediated by elevated levels of allostatic load during childhood. Employing an EF battery consisting of working memory, inhibitory control, and attentional control tasks, Blair and colleagues (2011) showed that poverty early in life (7–24 months) is related prospectively to worse EF skills at 3 years of age. Moreover, this relation was partially mediated by elevated cortisol levels during the same period of poverty. Recently, Sarsour and colleagues (2011) uncovered evidence that negative relations between family SES and 8- to 12-year-olds' performance on indices of working memory were partially mediated by cognitive enrichment in the home whereas Stroop performance was mediated by parental responsiveness and family companionship. Finally, Doan, Fuller-Rowell, and Evans (2012) examined in a prospective, longitudinal design an index of cumulative risk that included multiple factors, such as poverty, single-parent status, high school dropout, family turmoil, and violence in relation to internalizing and externalizing symptoms in late adolescents (16- to 18-year-olds). The adverse effects of cumulative risk on externalizing but not internalizing symptoms were mediated by a multimethodological index of self-regulation. Higher risk led to lower self-regulation skills, which, in turn, led to more behavioral conduct problems. Limited data suggest that links between SES and adverse developmental outcomes, including psychological distress and behavioral conduct disorders, may be explained, in part, by EF deficits. Examining heterogeneity in self-regulation skills, which includes both EF and emotion regulation, Garner and Spears (2000) found that more positive family expression of emotion along with parental consistency of discipline predicted better emotion regulation skills among low-income preschool children in their nursery school setting. Observers evaluated use of emotion regulation strategies, such as revenge, venting, or expression of dislike toward another child in response to anger and sadness. In a program of research on low- and working-class African American families in rural Georgia, Brody and his colleagues (e.g., Brody et al., 2006; 2013) were the first group to explore the potential role of EF and coping in children's mental health in relation to early deprivation. Household income among 9- to 12-year-olds in one sample was positively related to a standard index of children's self-regulatory abilities that largely tapped their planfulness and ability to reflect (“thinks ahead of time about the consequences of his or her actions”). Parents and teachers filled out this self-regulatory rating scale. Self-regulatory abilities, in turn, mediated adverse impacts of poverty on both internalizing and externalizing symptoms. Using similar measures, Brody and Flor (1997) showed that parents' perceived financial strain among the households of 6- to 9-year-old African American children was also associated with worse self-regulatory skills. As in the previous study, deficits in self-regulatory skill mediated the link between deprivation and

psychological well-being. In a second study focused on 7- to 15-year-old, low-income African American children living in single-parent families, Brody, Dorsey, Forehand, and Armistead (2002) replicated their earlier study plus found that more positive classroom processes, such as higher organization, better rule clarity, and high student involvement, along with more positive parenting processes, contributed to better self-regulatory skills, which, in turn, led to better child adjustment. Sektnan, McClelland, Acock, and Morrison (2010) found that parental ratings of self-regulatory behaviors (staying focused and on task plus ability to inhibit inappropriate responses during tasks) in their child at 54 months along with similar teacher ratings in kindergarten were significantly related to both family income and maternal educational attainment. The adverse impacts of low maternal education and low family income on math, reading, and vocabulary scores in first grade were partially mediated by selfregulatory behavior ratings at both developmental periods. More recently Crook and Evans (2014) have shown in a large heterogeneous, national sample that poverty during infancy predicts subsequent deficits in a planning task among preschool children. The Tower of Hanoi planning task requires participants to rearrange wooden rings onto a series of vertical poles from the ring that is smallest in diameter at the top to the ring that is largest diameter on the bottom of each pole. Only one ring at a time can be transferred from one pole to another. Performance on this planning task, in turn, was linked in a prospective, longitudinal analysis to fifth-grade math and reading standardized test scores. IQ was statistically controlled throughout. A smaller number of studies show that enhanced EF can serve as a resource enabling low-SES children to have better mental health outcomes than might otherwise be expected. Lengua, Bush, Long, Kovacs, and Trancik (2008) showed that the adverse mental health sequelae of poverty among third to fifth graders were attenuated among children with better attentional control. The EF index consisted of a composite of two different behavioral assessments of EF: the Stroop task and a Simon-says task, where the rules are periodically reversed. In a similar study, Veenstra, Lindenberg, Oldehinkel, DeWinter, and Ormel (2006) found that preadolescents from low-SES backgrounds engaged in more antisocial behaviors. This welldocumented relation, however, was significantly attenuated among those with high effortful control, based on a parental rating scale. Buckner and colleagues (2003) were interested in whether resilience among low-income children and youth (8–17 years) would be influenced, in part, by better EF skills. They used a Q sort rating conducted by highly trained clinicians who were quite familiar with each participant in order to quantify EF (e.g., “Is attentive and able to concentrate”). Resilience was defined in terms of good mental health, academic performance, and overall competency as assessed on a series of standard instruments. Low-income youth who were resilient had significantly higher EF scores. Furthermore, higher EF scores predicted subsequent socio-emotional and academic outcomes over a 2-year period (Buckner, Mezzacappa, & Beardslee, 2009). Low- relative to middle-SES children tend to have compromised EF—EF is a fundamental resource that aids in the regulation and management of stress. The direct, adverse impacts of chronic stress related to impoverished environments likely account for some of the covariation between SES and EF. A few studies also suggest that EF itself can mediate some of the mental

health sequelae of childhood poverty. For older children, EF may function more commonly as a moderator, providing some modicum of protection against the ill effects of deprivation on children's mental health. Just as self-regulatory skills can help us understand some of ways in which a low-SES childhood can lead to psychopathology, there is also evidence showing that coping strategies can alter the impacts of low SES on mental health outcomes. One noticeable omission in the SES and coping research is a paucity of research comparing coping strategies between lowand higher-SES children or youth. Neuendorf, Kim, and Evans (2009) found in two different studies that low- relative to middle-income adolescents (age 16–18) were more likely to rely on disengagement coping strategies, such as denial and avoidance, rather than on engagement coping strategies, like problem solving or marshaling social support, when dealing with an important, recent stressor in their life. Most coping and SES studies have examined linkages between different coping strategies and psychological health within low-SES samples. A smaller number of researchers have also investigated whether coping strategies help explain the relation between poverty-related stress and mental health outcomes. Low-income children in fourth and sixth grade, who are more resilient in terms of good behavioral adjustment evidenced more effective interpersonal problem-solving skills and relied more on positive coping skills, such as self-reliance and support seeking, relative to their nonresilient peers (Parker, Cowen, Work, & Wyman, 1990). Tolan, Gorman-Smith, Henry, Chung, and Hunt (1998) investigated the relations between various coping strategies and mental health outcomes in a sample of low-income, predominantly African American and Hispanic adolescents (age 12–16) living in the inner city. Cross-sectionally, youth who relied predominantly on problem solving and quiet substance users (reliant on avoidance through substance abuse but overall low levels of coping) had fewer internalizing symptoms. At the same time, cross-sectional data also revealed that emotion-focused coping as well as youth who depended heavily on avoidance through substance abuse but who also vented emotionally (emotional substance users) had higher levels of externalizing symptoms. One year later both the quiet and the emotion-focused substance abuse users showed the greatest rise in externalizing symptoms. The quiet substance users also showed a significantly greater rise in internalizing symptoms. Gonzales, Tein, Sandler, and Friedman (2001), looking at a younger adolescent sample (seventh–eighth grade) in similar circumstances (poor, inner city, predominantly non-White), found that active coping and distraction, respectively, buffered the relation between family stress and conduct problems in girls; active coping did the same for depression in boys. These data on coping and psychological well-being within low-SES samples are consistent with Wadsworth's program of research on poverty-related stress, coping, and mental health (Wolff et al., 2011). Adolescents from low-income families who cope with poverty-related stressors using engagement coping strategies (e.g., problem solving, emotion regulation, distraction, cognitive restructuring) tend to fare better psychologically than their low-income peers who instead rely more on disengagement strategies, such as avoidance or denial (Wadsworth & Berger, 2006; Wadsworth & Compas, 2002; Wadsworth et al., 2005; Wadsworth & Santiago, 2008).

Further analyses also suggest that some of the ill effects of poverty-related stress on internalizing and externalizing symptoms among adolescents are partially mediated by coping strategies with engagement strategies, particularly secondary control techniques, such as cognitive restructuring and distraction, which are helpful in attenuating negative impacts (Wadsworth & Compas, 2002; Wadsworth et al., 2005). Interestingly, in the Wadsworth et al. (2005) study, for parents, more engagement buffered rather than mediated the negative correlates of poverty-related stress on their own mental health outcomes. For parents, disengagement strategies had the opposite effect, accentuating the negative relations between poverty-related stress and psychological distress. The findings of coping strategy mediation among adolescents and the interaction between coping strategies and poverty stress among adults suggests the possibility of a developmental sequence with coping strategies of older individuals being less malleable in comparison to those of children and adolescents. However, in a prospective, longitudinal study of low-income middle and senior high students, Wadsworth and Berger (2006) found one interaction between poverty stress and primary control strategies. Youth higher in poverty stress had greater increases in internalizing symptoms over an 8-month period, but this was largely attenuated if they relied primarily on primary control coping strategies. This interaction between poverty-related stress and primary control coping, however, was the only moderation detected among several possible interactions between coping strategies and poverty stress. Gonzales and colleagues (2011), however, did detect an interaction between various sources of stress (family, peer, community) and coping on mental health outcomes within their low-SES middle school sample. Both active coping and distraction buffered the adverse effects of family stress on behavioral conduct for girls. For boys, the effects of peer stress and community stress, respectively, on depression were exacerbated by more active coping. The authors suggest that boys who try to actively cope with intractable, chronic stressors like peer or community violence may suffer the frustration of invoking an instrumental coping strategy when in fact accommodation might fit better with the nature of the demands. One possible reason for the discrepancy between the absence of coping moderator effects in the Wadsworth and Berger (2006) study compared to that of Gonzales et al. is that Wadsworth and Berger examined stress directly related to poverty (not enough money to buy clothes, parents worried about paying bills), whereas Gonzales et al. examined various sources of stress (peers, community, family) in conjunction with coping strategies. Edith Chen and Greg Miller have been conducting a program of research examining how cognitive appraisals and adjustments to the stressors accompanying low-SES childhoods can affect health and well-being. In an initial series of studies, youth who saw videotapes of simulated encounters between adolescents evaluated how threatening/hostile the situations were. Low-SES youth were more likely to attribute threat or hostile intent in ambiguous social interactions compared to their middle-SES peers. These attributions, in turn, were linked to elevated physiological stress responses (Chen & Matthews, 2001; Chen, Langer, Raphaelson, & Matthews, 2004). More recently, Chen and Miller (2012) initiated a series of studies examining a coping strategy, shift and persist, as a potential protective process among lowincome children. Shifting refers to the ability to adjust individual responses to stressors

through emotion regulation strategies, such as cognitive restructuring. Persisting reflects enduring adversity by strategies such as finding meaning in difficult circumstances or remaining optimistic. Asthma occurs more often and with greater severity among low-income children. In a sample of children (mean age 12) with diagnosed asthma, low-SES children who were high in shift and persist had lower levels of inflammation and degrees of impairment (e.g., school absence) compared to their low-SES peers with low shift and persist skills (Chen, Strunk, et al., 2011). In fact, low-SES asthmatic children who used this coping strategy resembled their high-SES counterparts. In a different sample, childhood SES was inversely related to allostatic load in a national sample of middle-age adults (Chen, Miller, Lachman, Gruenewald, & Seeman, 2012). However, adults from more deprived backgrounds who engaged in more shift and persist did not have higher allostatic load scores than their counterparts from higher SES backgrounds. Another coping resource that appears capable of buffering some of the ill effects of childhood poverty on mental health is social support. As noted earlier, one reason for this could be because support appears to reduce the negative effects of stress on parenting behaviors of lowincome parents. A large developmental literature demonstrates the protective effects of maternal responsiveness on socio-emotional development among children (Bornstein, 1989; Bradley, Corwyn, Burchinal, McAdoo, & Garcia-Coll, 2001; Bradley, Corwyn, McAddo, & Garcia-Coll, 2001; Shonkoff & Phillips, 2000). Studies of resilient children in the context of poverty routinely find that a strong, positive bond with an adult caregiver is among the most protective factors (Werner & Smith, 1982; Wyman, Cowen, Work, & Parker, 1991). Maternal responsiveness attenuates several adverse effects of childhood deprivation, including elevated levels of allostatic load in adolescents (Evans, Kim, Ting, Tesser, & Shanis, 2007), the metabolic syndrome in middle-age adults (a cluster of risk factors predictive of diabetes, stroke, and cardiovascular diseases) (Miller et al., 2011), and proinflammatory processes among healthy 25- to 40-year-olds (Chen, Miller, Kobor, & Cole, 2011). There could be a link between responsive parenting and the development of children's own adaptive coping responses. Children who grew up in families with multiple problems, including poverty, but who had more responsive and nurturing parents develop enhanced self-regulatory skills (Repetti et al., 2002). They may also learn that enlisting support from others is effective. Enlisting such support may be particularly important given the well-documented inverse relation between SES and social support (Evans, 2004; Matthews & Gallo, 2010). Low-SES children and their parents typically have fewer social resources they can call on. Both EF and coping are central psychological processes involved in the relations between SES and child psychopathology. A robust finding is that children and adolescents who cope with poverty-related stress by relying primarily on disengagement strategies, such as avoidance or denial, have worse mental health outcomes. Problem solving, emotion regulation, finding meaning, and maintaining optimism, along with the recruitment of social support, appear to be adaptive strategies that help counter some of the ill effects of impoverishment in regard to mental health.

Brain and SES

An exciting, emerging area of scholarship in the neurosciences is examining the potential role of the brain in linking SES and other forms of early childhood deprivation to psychopathology as well as cognitive outcomes (Gianaros & Manuck, 2010; Hackman & Farah, 2009; Hackman, Farah, & Meaney, 2010; McEwen, 2012; McEwen & Gianaros, 2010; Raizada & Kishiyama, 2010). Although the focus of much of this work is on exploration of cortical and subcortical mechanisms responsible for the SES-achievement gap, a small amount of the research has potentially major implications for the development of psychopathology. Much of the initial neuroscience work examined neurotransmitter processes or conducted electrophysiological protocols, but recently the field has moved more into structural and functional neuroimaging techniques. Attenuated serotonin responsivity is related to elevated symptoms of aggression as well as depression. Lower-SES adults have attenuated serotonin responsiveness (Manuck et al., 2005; Matthews, Flory, Muldoon, & Manuck, 2000). There may also be a genetic component of the serotonin neurotransmitter system and the effects of early life deprivation. The positive association between SES and serotonin responsiveness is accentuated among persons with the short allele on the 5-HTTLPR serotonin transporter gene (Manuck, Flory, Ferrell, & Muldoon, 2004). As we discuss later, this genetic polymorphism may be a salient marker for vulnerability to chronic stress. A small number of electrophysiological brain wave recording studies suggest that children from lower SES backgrounds have delays in development of the prefrontal cortex (PFC), which is a major cortical area responsible for executive functioning, such as the self-regulation of behavior and emotional control (Heatherton, 2011). Earlier we reviewed several studies indicating that childhood SES is positively associated with better executive functioning. The PFC is believed to operate as a top-down regulator of behavior enabling individuals to inhibit strong emotional responses and to reflect on various coping strategies in contemplation of action. Examining brain wave activity at 18, 30, and 48 months, Otero (1994, 1997) found evidence of delayed maturation in the prefrontal PFC of children from low-SES families relative to children from middle-SES families. Early extrastriate attention-sensitive components and novelty-evoked related potential (ERPs) are attenuated in low-SES relative to high-SES 7- to 12-year-olds who also had more difficulties with several neurocognitive measures of executive functioning (Kishiyama, Boyce, Jimenez, Perry, & Knight, 2008). Threeto 8-year-olds and middle school children from lower-SES compared to middle-SES families appeared to require greater effort to selectively attend to auditory stimuli while gating out distracting background stimuli (D'Angiulli, Herdman, Stapells, & Hertzman, 2008; D'Angiulli, Weinberg, Grunau, Hertzman, & Grebenkov, 2010; Stevens, Lauinger, & Neville, 2009). The positive indication from electroencephalographic research that low-SES children have depleted attentional capacity fits well with several of the EF studies reviewed earlier that showed compromised attentional control abilities among low-SES children. Evidence from SES-related brain structural and functional imaging studies also implicates childhood SES in brain development. Similar to the electrophysiological studies on SES and brain activity, Sheridan, Sarsour, Jutte, D'Esposito, and Boyce (2012) found that 8- to 12-yearolds from lower-SES families had more difficulty learning a response discrimination task that appeared to be related to greater activation in a portion of the PFC, the right middle frontal

gyrus. Increased neural recruitment in this area may have reflected greater effort or less efficiency in the learning of the rules necessary to perform optimally on the task. In a series of studies, Gianaros and colleagues (2007) have shown both structural and functional alterations in brain development among low-SES relative to middle-SES individuals with potentially profound mental health implications. Adults who perceive themselves as lower in relative social status have lower gray matter volume in the anterior cingulate cortex, a limbic brain area centrally involved in emotional regulation as well as responsivity to stressors. Similar volumetric reductions in the anterior cingulate cortex are associated with mood disorders and depression. As indicated, self-regulatory and executive control functions appear to be damaged in adults and children from low-SES childhoods. Adults from low-SES childhoods evidenced reduced connectivity between portions of the PFC and orbitofrontal and striatial areas of the brain implicated in reward sensitivity and impulse control (Gianaros et al., 2010). In a second study, adults who grew up in families with perceived lower relative social standing had elevated amygdala activation to emotionally threatening faces but responded similarly to their peers from family backgrounds of average and above average social standing to neutral and nonthreatening faces (Gianaros et al., 2008). Similarly, adolescents from lower-SES families evidenced greater amygdala as well as dorsomedial PFC activation in response to emotionally threatening faces (Muscatell et al., 2012). Yanagisawa and colleagues (2012) asked college graduates how much financial strain their family faced when they were growing up. Higher levels of perceived financial strain were associated with greater social distress and lower levels of activity in the right ventrolateral PFC during a social exclusion task. In the cyberball task, participants play a virtual game of catch with what they believe are two other players (whose pictures are displayed during the game); in actuality, participants are interfaced with a computer that exposes participants to two different conditions: inclusion in the game of catch and exclusion in the game of catch. The former serves as a baseline contrast for the exclusion phases of the experiment. The pattern of findings from these recent neuroimaging studies on greater threat activation when confronted with emotionally charged stimuli fits nicely with the work by Chen and colleagues, reviewed earlier, showing that low-income youth more readily attribute threat and hostile intent to ambiguous interpersonal encounters (Chen & Matthews, 2001; Chen et al., 2004). Another area of the brain, the hippocampus, helps integrate information about emotionally laden information to past experiences and context, better enabling the organism to respond more adaptively to challenging situations. Considerable animal work indicates damage to the hippocampus at the cellular level from chronic stress as well as reduction in gray matter volume (Lupien et al., 2009; McEwen & Gianaros, 2010). Recently different teams of investigators have linked lower-SES childhoods to smaller hippocampal volumes in 5- to 17year-olds (Noble, Houston, Kan, & Sowell, 2012), in 9- 11-year-olds (Hanson, Chandra, Wolfe, & Pollak, 2011), and in both middle-age (Butterworth, Cherbuin, Sachdev, & Antsey, 2012) and older adults (Staff et al., 2012). Recently, Pilyoung Kim and her colleagues (Kim et al., 2013) integrated these brain and self-regulatory behavior findings with the findings from stress research, finding that adults who grew up in poverty had reduced activation in the PFC

and less efficient regulation of the amygdala during a task in which they had to regulate negative emotions. Furthermore, chronic stressor exposure during childhood mediated the association between poverty and PFC activity. Taken in concert, this newly emerging literature on SES and the brain appears to hold great promise for unpacking the underlying biological mechanisms that help transmit the experience of childhood poverty and deprivation into adverse mental health outcomes. In particular, the neuroscience findings suggest that persons from low-SES backgrounds may respond more strongly to aversive and threatening stimuli in areas of the brain associated with autonomic, rapid responses to threat and emotionally negative stimuli (e.g., the amygdala). At the same time, persons from lower-SES backgrounds appear to have a less well developed top-down regulatory PFC system for controlling these rapid, automated emotion responses. Thus, individuals growing up in low-SES environments may be subject to a kind of triple jeopardy when it comes to responding to adverse experiences. First, they are likely to encounter many more psychosocial and physical risk factors that are threatening. Second, low-SES children appear to be downwardly regulated, being hyperresponsive or hypersensitive to threatening events. Third, at the same time, they are less able to monitor, regulate, and modify these rapid, emotion-laden responses to threat and other negative emotionally evocative situations. One major shortcoming of all of the neuroscience work to date on SES is the absence of direct evidence for structural or functional changes in the brain operating as mediators or underlying explanatory mechanisms directly accounting for well-established linkages between SES and psychopathology. We currently do not have evidence for the full pathway: Low SES → Brain Alteration → Psychopathology.

Practice and Policy Community and/or Government Interventions: Policy Several safety net programs were designed by the federal government in an effort to meet the goals of the War on Poverty efforts of Presidents Kennedy and Johnson in the early 1960s. These programs, collectively referred to as social welfare, have parallels in most every industrialized nation, and are designed to prevent individuals and families from starvation and homelessness. As shown in national poverty rate trends, programs such as Aid to Families with Dependent Children (AFDC, formerly known as Welfare, and now referred to as Temporary Assistance to Needy Families, TANF), WIC, and SNAP contributed to significant reductions in poverty rates, from a high of 22.4% in 1959 to a low of 11.1% in 1973. Unfortunately, despite continued increases in the GDP since the 1950s, the poverty rate has not seen further reductions since the early 1970s, in part because eligibility levels for these programs have not been adjusted for inflation and cost-of-living increases over the last four decades. Such safety net programs have also been documented to produce improvements on a variety of child health and developmental outcomes, such as reduced low-weight infant births (Kowaleski-Jones & Duncan, 2000); lower risk of abuse and neglect; and fewer nutritional deficiencies, such as anemia and failure to thrive (Lee & Mackey-Bilaver, 2007).

In 1975, the Earned Income Tax Credit (EITC) was passed to benefit low- and moderateincome working families, designed to offset the burden of Social Security taxes and to provide an incentive to work. The EITC was also successful in raising the incomes of working families, and Dahl and Lochner (2012) found that each $1,000 increase in family income stemming from the EITC was accompanied by significant increases in math and reading achievement, especially for children in the most disadvantaged families. Similarly, Strully, Rehkopf, and Xuanc (2010) found that state EITCs led to increased birthweights and reduced maternal smoking in families in poverty. The estimate of the $500 billion annual cost of childhood poverty to the United States calculated by Holzer et al. (2008) suggests that there are large potential gains to be had by poverty-reduction programs and other types of interventions targeting the environment of childhood poverty. Holzer et al. outlined the next types of programs as having the potential to reduce poverty and have high monetary returns on the investment: programs that increase wages and incomes (e.g., increases in minimum wage, EITC), programs that increase earning potential via training and education, programs that improve neighborhood conditions and housing quality, and universal high-quality preschool programs. Calculations by Dickens, Sawhill, and Tebes (2006) suggested, for example, that making high-quality preschool programs universally available to children could increase GDP by 3.7% annually, coming close to offsetting the $500 billion price tag of child poverty in the United States. Knudsen, Heckman, Cameron, and Shonkoff (2006) similarly reviewed the range of potential benefits to be gained from early intervention programs for disadvantaged children, concluding that the estimated return on each dollar invested in early childhood programs, such as the Perry Preschool Program, exceeded 17%, which these authors noted is a much higher rate of return than standard stock market returns. With positive outcomes in adulthood spanning academic achievement and attainment, cumulative wage earnings, likelihood of owning a home, likelihood of receiving public assistance, and years spent incarcerated, early childhood interventions have a very large potential to translate into high gains in human and economic capital. Heckman (2006) estimated that the total cost of implementing the Perry Preschool Program was $16,514 (in 2004 dollars) per child and that benefits saved by implementation across earnings, special education, crime, and public assistance totaled $144,345, with a benefit to cost ratio of 8.74. The results of Nuru-Jeter et al.'s (2010) study of categorical versus continuous measures of SES and relations with mental and physical health outcomes suggest that “efforts to alter the social conditions that drive health effects will demand mobility into meaningful strata of educational and financial well-being” (p. 17). Hence, while reducing low birthweights and reducing hunger are incredibly important and worthy goals, efforts to improve poor children's life chances must do more than marginally increase family incomes; rather, they must strive to help children achieve the education and training they need to rise in social stature. Consistent with the message of Nuru-Jeter et al. (2010), Zaslow et al. (2002) reviewed the impacts of four kinds of welfare-reform programs on child outcomes: (1) work-first or education-first programs that make assistance contingent on getting a job or job training (e.g., Job Opportunities and Basic Skills Training); (2) imposing time limits on public assistance

receipt (e.g., Florida Family Transition Program); (3) programming delivered specifically to teenage mothers (e.g., New Chance Demonstration); and (4) financial incentives tied to working (e.g., New Hope Project). Of the four program types, Zaslow et al. found that only the programs that led to sustained financial gains for families had favorable impacts on child health and well-being. Programs that failed to increase employment and income had either neutral or detrimental effects. Similarly, in analyses making use of random-assignment experimental design and instrumental variables estimation, Gennetian and Morris (2000) found evidence that increasing income via the Minnesota Family Investment program can improve children's school engagement and social-emotional functioning. Their findings further supported the conclusion that income has a causal and reversible impact on child functioning. A new generation of antipoverty programs has expanded on this strategy of contingent safety net provision by making public assistance contingent on poor parents performing a number of stipulated tasks designed to increase their children's human capital. Such conditional cash transfer (CCT) programs reward parents with money for taking their children to medical checkups, getting them vaccinated, and improving school attendance. Some CCT programs implemented outside the United States have demonstrated decreased rates of poverty and child illness and improved school attendance (e.g., Rivera, Sotres-Alvarez, Habicht, Shamah, & Villalpando, 2004). Fernald, Gertler, and Neufeld (2008), for example, examined child outcomes associated with the Mexican CCT program Opportunidades. A doubling of income via cash transfers was associated with better infant and toddler outcomes, such as taller height for age, less stunting, lower body mass indexes as well as higher motor, cognitive, and language development scores. CCT programs have been implemented in several other lowerand middle-income countries, including Brazil, Peru, Honduras, Nicaragua, Colombia, India, and Nepal. Overall, outcomes are mixed (Wolf, Aber, & Morris, 2013), with short-term positive effects found in some countries, but not others (e.g., Fiszbein & Schady, 2009). Only one large-scale CCT program has been evaluated in the United States (Opportunity New York; Riccio et al., 2010). Results of this evaluation showed that, while the program decreased current poverty and material hardship, it did not result in improved school outcomes for elementary or middle school students. It did however, lead to improved academic outcomes for high school students who were already meeting grade-level standards at the beginning of the study.

Implications of Poverty and Psychology Research for Policy Not surprisingly, U.S. government programs that directly increase the income of families living in poverty have reduced poverty rates. What is surprising is that Americans have allowed these programs to lose power by not indexing them for inflation. Research on the effects of poverty on children has provided more data than needed to justify increased investment in antipoverty programs; yet the public, in general, has not prioritized such expenditures—despite the fact that most, if not all, Americans believe in an ethic of justice (Hamlin, Wynn, Bloom, & Mahajan, 2011; Masters & Gruter, 1992). The disconnect between Americans' stated values and voting patterns suggests that developmental psychopathologists who are committed to reducing the negative effects of

poverty on children might consider bringing psychological research to bear on factors affecting U.S. voting priorities. Psychological constructs, such as cognitive dissonance (Cooper, 2007; Festinger, 1957/1985) and belief in a just world (Furnham, 2003), can help explain why voters may blame others for negative experiences that would otherwise simply seem unfair (Godfrey & Lowe, 1975; Janoff-Bulman, Timko, & Carli, 1985). In the case of cognitive dissonance, for example, middle- and upper-class individuals are likely to experience discomfort associated with incongruous beliefs, such as: (a) there are children in our country who are living under life-threatening conditions of violence because they reside in urban poverty; (b) we live our lives without substantially addressing this problem; and (c) we are good, kind people. The just world belief (i.e., that those who suffer have brought that suffering upon themselves in some way; Lerner, 1980; Montada & Lerner, 1998) can reduce that cognitive dissonance. If those who are disadvantaged deserve their fate, then the rest of us can bear less guilt for not intervening (Lodewijkx, de Kwaadsteniet, & Nijstad, 2005; Loseman & van den Bos, 2012). We can eliminate the discomfort of the cognitive dissonance associated with the three statements by adding a fourth: Everyone who works hard can escape poverty and violence (Murray, Spadafore, & McIntosh, 2005; van den Bos & Maas, 2009). Two additional psychological constructs may help explain the lack of investment in addressing poverty. The first is diffusion of responsibility (Darley & Latané, 1968; Latané & Darley, 1968). If one believes that action does not rely solely on oneself, one is less likely to act (Leary & Forsyth, 1987; Mynatt & Sherman, 1975). The second is a lack of knowledge of how one can effectively intervene (Fielding & Head, 2012). When pathways to successful intervention are unclear and the outcome is uncertain, individuals are less likely to act (Jorm, 2012). Most developmental psychopathologists who conduct poverty research have focused on poverty's effects on children's mental health and the mechanisms of those effects. This work lays the groundwork for psychological interventions (as summarized in the next section). Given that a lack of knowledge of how to intervene effectively reduces the likelihood that individuals will intervene, developing effective interventions also is important for policy efforts. Furthermore, despite the fact that safety net programs can protect children from some of the most dire effects of poverty (malnutrition, homelessness), they have not proven sufficient to alter the economic trajectories of children born into poverty (or poverty would have been eradicated by now). And welfare-reform programs designed to alter those trajectories by bolstering the capacity of low-income families to increase their own incomes (through education, training, or investment in the training and education of their children) have had only mixed effects, as reviewed earlier. Thus, there is a real need for the development of effective programs that can change the economic trajectory of children born into poverty. There is strong evidence that changing children's educational trajectories can change their economic trajectories and break the cycle of poverty (Haskins & Rouse, 2013). Since the 1980s, earnings have increased only for families that include an adult with a college degree (Haskins & Rouse). As a result, policy experts like Haskins and Rouse have proposed that federal programs focus on increasing the preparedness of low-income students for college.

Early childhood programs like Head Start have demonstrated success, but they have not succeeded in eliminating achievement or income gaps for low-income children, and programs that target older children for college have had only modest effects at best (Haskins & Rouse, 2013). So there is great need for additional intervention development and evaluation efforts. To coordinate and support such efforts, Haskins and Rouse recommend that the federal government (a) consolidate existing funding efforts into a single grant program that requires funded initiatives to be backed by rigorous research and (b) give the U.S. Department of Education the authority to plan and execute a coordinated set of studies that develop and rigorously test several approaches to college preparation. A team of scholars and practitioners in Chicago is following the recommendation to focus on college preparedness while also addressing barriers to middle and upper-class investment in anti-poverty programs. The Chicago Public School System (CPS) is 87% low income, and only 8% of its students graduate from a 4-year college (Allensworth, 2006; Chicago Public Schools, n.d.). With the goal of changing those statistics, CPS has partnered with five major universities in the city of Chicago—DePaul University, Loyola University, Northwestern University, the University of Chicago, and the University of Illinois at Chicago—to submit a proposal to the U.S. Department of Education that is focused on both (a) supporting the academic preparedness of CPS students for college and (b) increasing the sense of responsibility for ending poverty among college students through an innovative program called Cities Mentor Project. Cities Mentor Project provides early adolescent CPS students with (a) training in researchbased strategies for engaging with academic challenges and coping with severe and chronic stressors endemic to urban poverty that impede learning (e.g., community violence); (b) connection to undergraduate mentors who support youth engagement and coping efforts in reallife situations, advocate for youth academically, and connect youth to (c) high-quality afterschool and summer programming that provides additional support. Cities Mentor Project makes a long-term commitment to youth, supporting them from middle school all the way through their first year in college. University student mentors in the program take part in a service learning course focused on training and supporting their mentoring work. The class also educates the mentors about the roots and effects of poverty in Chicago and brings the powerful cognitive-behavioral tools of Socratic questioning and collaborative empiricism (which are typically used to target mental health problems; Cohen, Edmunds, Brodman, Benjamin, & Kendall, 2012; Tee & Kazantzis, 2011) to bear on psychological barriers to moral action (Hart et al., 2009; Overholser, 2011; Palmer & Williams, 2013). Sample questions used in the class include: (a) What is the cause of poverty and inequality in the city of Chicago? (b) Who is to blame? (c) What can be done about it? (d) What should I do about it as an individual? (e) What should we do about it as a group? Thus, this program moves beyond intervention strategies characteristic of traditional psychology to include community psychology principles of empowerment and advocacy, principles that may be necessary to change policy (Dalton, Hill, Thomas, & Kloos, 2013; Speer, Peterson, Armstead, & Allen, 2013).

The CPS-university partnership plans to evaluate the effects of this new approach on a broad range of outcomes, for mentors and mentees alike, including mental health and academic achievement as well as moral understanding and civic engagement. If there are expected empirical changes on these measures, this partnership could become a model not only for changing the educational and economic trajectories of low-income children but also for increasing a sense of responsibility for ending poverty among more highly resourced individuals. Based on our review of research and policy, it appears we need both.

Individual and Family Interventions: Practice Numerous early intervention programs have been developed with the aim of promoting educational and socio-emotional success and preventing dysfunction stemming from poverty and socioeconomic disadvantage in young children. In a comprehensive review of results from 20 rigorously evaluated programs, including a mix of home visiting, parent education, early childhood education (ECE), and combination programs (e.g., Abecedarian Project, NurseFamily Partnership), Karoly and colleagues (2005) report positive benefits, both immediately following intervention (e.g., IQ scores, prosocial outcomes) and at long-term follow-up (e.g., educational attainment, teen pregnancy, arrests) to participating children. Successful early childhood intervention programs tend to have multiple components that target parents (e.g., job-related services), children (e.g., enrollment in high-quality ECE), and/or the parent-child dyad (e.g., parenting classes). Karoly and colleagues also identified several program elements that characterized effective programs, including well-trained program leaders (e.g., home visitation services provided by trained nurses), intensive program scope and delivery, and center-based education-focused child care in settings with small child-staff ratios and welleducated lead teachers. Several successful large-scale programs have been in operation for several decades, allowing for evaluation of the long-term outcomes possible with high-quality ECE. Head Start (e.g., Ripple & Zigler, 2003) and the Perry Preschool Program (Weikart & Schmeinhart, 1992) are successful prevention programs targeting poor families with preschool children designed to prevent the emergence of behavioral difficulties and promote positive social skills, emotional competence, and academic achievement. Both of these programs focus on providing families with access to basic human necessities, such as food, housing, education and health care, as well as high-quality child care. These programs have produced both substantial initial effects on child IQ scores and social/emotional outcomes as well as employment, criminality, and family stability outcomes that persist until well into midadulthood (e.g., Pungello et al., 2010; Ramey & Ramey, 1998). Brooks-Gunn (2003) challenged us to find any other interventions that come close to the magnitude of these kinds of effects in the learning sciences (and beyond). The Nurse-Family Partnership program has produced strong outcomes for families who receive perinatal and infancy home visits from qualified nurses (e.g., Olds, Sadler, & Kitzman, 2007). A recent evaluation by Eckenrode and colleagues (2010), for example, examined results of the program on the life course development of 19-year-olds whose mothers had participated in the program. Girls from the home visitation group were significantly less likely to be arrested (10% versus 30%) and convicted of a crime (4% versus 20%) and had fewer lifetime

arrests (.10 versus .54) than girls in the control group. Additionally, for high-risk girls in particular, significantly lower rates of teen pregnancy and public assistance usage were found for home visitation girls in comparison to controls. The Strong African American Families program was developed by Gene Brody and colleagues (2006) to prevent substance use, conduct problems, and depression in rural, low-income African American adolescents. This program, which built on over a decade of basic research on rural African American families and their strengths, focuses on building caregiving practices, such as limit setting, monitoring, instilling racial pride, and effective problem solving. Adolescent self-regulation is also directly targeted. Participation in the 5-session intervention was associated with a 36% reduction in the frequency of conduct problems, a 32% decrease in overall substance use, and a 47% decrease in substance use problems over the course of almost 2 years (Brody et al., 2012). The Families Coping with Economic Strain intervention (FaCES; Raviv & Wadsworth, 2010) teaches children (age 8–12) and their parents coping strategies to effectively manage povertyrelated stress. Sessions introduce various ways of coping, teach emotion regulation skills, and educate parents and children about the effects of poverty on the family and the benefits of having strategies to utilize when stressors occur. An evaluation of FaCES using a multiple baseline design, found that children increased their use of effective coping, which was accompanied by substantial decreases in internalizing and externalizing symptoms (Raviv & Wadsworth, 2010). Parenting-specific interventions have also been shown to improve dysfunctional parenting styles by addressing the parenting practices involving appropriate discipline, promotion of prosocial behaviors, and engagement with children's schools (e.g., Brotman et al., 2011). For example, a new family-strengthening intervention developed to address the key parts of family stress model targets strengthening the interparental relationship, improving coping and reducing stress, and teaching child-centered parenting. This intervention has demonstrated improvements on parenting, parental coping, and depression that appear to translate into improved child internalizing and externalizing symptoms (Wadsworth et al., 2013). Similarly, the family-focused SAFEChildren (Schools and Families Educating Children) intervention targets supportive parenting, teaches parental skills for managing neighborhood challenges, and promotes school involvement among low-income parents. The program includes a tutoring component to improve child academic achievement. Program evaluation results have shown significant improvements in parents' use of stable, consistent caregiving and limit setting as well as improvements in children's behavior and attention regulation (Tolan, Gorman-Smith, Henry, & Schoeny, 2009). Finally, some school-based interventions targeted to school-age children have also shown promise. The Fast-Track program, for example, provides a wide array of basic services to families. In addition, however, Fast-Track incorporates the teaching of basic self-regulation skills into units developed for both parents and children. Children participating in Fast-Track receive a classroom program, called PATHS (Promoting Alternative Thinking Strategies). PATHS helps children work on self-control skills by teaching emotional awareness and

understanding. It also helps children enhance peer-related social skills and social problem solving to increase social competence (Bierman, 2002). The program, which began with a sample of first graders and followed them through the end of elementary school, has been successful in reducing the incidence of home and community conduct problems, involvement in peer deviance, and social incompetence (Dodge & Conduct Problems Prevention Group, 2007).

Implications of Poverty and Psychology Research for Practice Studies on mediators and moderators of the effects of poverty are especially relevant for psychological practice focused on mitigating the effects of poverty on children's mental health, as these studies highlight mechanisms that can be targeted for intervention and indicate groups and contexts that could influence efficacy. Implications of findings from each of these literatures are summarized in the next sections.

Implications of Mediator Data for Future Interventions Individual and Family Mediators Basic research on mediational processes linking poverty exposure to negative outcomes is especially relevant to intervention, as findings in this area highlight pathways between poverty and symptoms that can serve as intervention targets. To date, mental health interventions have built primarily on mediational findings related to parenting and stress exposure. In particular, as outlined earlier, interventions have focused on teaching parents and children to manage poverty-related stressors in ways that support healthy family functioning and disrupt pathways linking poverty to youth psychological problems. Although coping, per se, has yet to emerge consistently as a significant mediator or moderator of poverty effects on youth mental health (Grant et al., 2006), the preliminary success of interventions targeting coping provides evidence that coping is an important intervention target (Raviv & Wadsworth, 2010). The constructs of self-regulation, executive functioning, and mastery are all interconnected with the construct of coping. In particular, coping interventions focused on helping youth identify which coping strategies to use in a given situation require use of self-regulation and executive functioning and, if effective, are likely to build mastery and reduce helplessness. As noted, a solid body of research has established various components of executive functioning as mediators of the negative effects of poverty. Nonetheless, few intervention research studies have examined changes in mental health as a function of changes in executive functioning. There also have been few studies to examine changes in self-regulation or executive functioning or mastery (beyond coping mastery) as a result of coping training or vice versa. A larger body of research has addressed mediation effects of executive functioning than of coping. Nevertheless, a greater number of mental health interventions have been developed to target coping. This phenomenon may reflect a belief among developmental psychopathology researchers that executive functioning interventions target academic outcomes and coping interventions target psychological outcomes. There is mounting evidence, however, that

executive functioning is an important prerequisite of effective coping, especially in light of findings that selecting the coping strategy that best fits a particular type of problem is essential for coping effectiveness (Compas, Connor-Smith, Saltzman, Thomsen, & Wadsworth, 2002; Compas et al., 2010). Additional intervention work is needed to better understand how executive functioning, selfregulation, mastery, and coping might be targeted synergistically. For example, it would be helpful to know whether there are benefits to targeting self-regulation or executive functioning in advance of coping training or whether coping training itself is the most efficient and relevant approach to teaching self-regulation and executive functioning in the context of poverty-related stressors. In addition, it would be ideal to integrate such intervention work with basic research by testing differences in basic cognitive functioning and brain structure as a result of coping training. Such research also is needed to help us understand how structural, physiological, cognitive, and emotional processes function together to mediate poverty. For example, the next questions remain: (1) Does poverty directly affect brain structures (e.g., through malnutrition), thereby reducing capacity to learn adaptive strategies, or do limited learning opportunities lead to reduced brain growth, or do both processes occur and affect one another reciprocally? (2) Which specific poverty-related stressors predict which mediating processes and disrupt which protective moderators? (3) In the case of multiple mediators (e.g., cortisol, cognitions, emotions, executive functioning), does one predict the other, or do they reflect different measures of the same process? Answering these questions through basic and/or intervention research will help determine intervention approaches most effective at disrupting naturally occurring pathways between poverty and negative outcomes. Peer, Neighborhood, and Physical Environment Mediators Far fewer poverty interventions focused on mental health outcomes have targeted neighborhood, physical environment, or peer mediators. This fact likely reflects the fact that targeting such variables requires considerably more resources than a typical mental health intervention and/or requires buy-in from large groups of individuals and/or requires shifts in social policy that fall beyond the purview of mental health providers. Nonetheless, a few housing studies have examined mental health effects of moving low-income families from lowincome, low-resource neighborhoods, characterized by community violence, poor physical environments, and high proportions of aggressive peers, to middle-class, high-resource neighborhoods with fewer risk factors (Fine, 2002; Leventhal & Newman, 2010). Findings for youth mental health have been mixed (Fine, 2002; Leventhal & Newman, 2010). One reason for the lack of consistent positive findings may be that youth who move from low- to highresourced communities experience negative effects associated with moving (e.g., disrupted social ties, being new to a school and community) and/or increased exposure to classism and racism. Much additional intervention research is needed to determine the most effective approaches for addressing neighborhood, peer, and physical environment mediators of the effects of poverty on mental health.

One potential approach would be to build on existing coping interventions as well as community psychology constructs of empowerment (Bond & Keys, 1993) to expand definitions of active coping to include activism. Current coping interventions that focus on matching particular coping strategies with particular stressors in the context of urban poverty teach youth to avoid community violence so that they do not (a) experience the internalizing distress that can result from trying to fix something beyond their capacity to fix or (b) display the externalizing behavior that may be required to stay safe under threatening circumstances (Grant et al., 2014). It might be useful, however, to teach youth that although community violence likely cannot be addressed by individually based active coping, it could be addressed by group-based active coping or activism. Such intervention would require partnerships between mental health providers and community-based organizations capable of involving youth in such work. Such interventions would also necessitate a lengthier investment than is typical of existing coping interventions (Gaylord-Harden et al., 2013). A recent meta-analysis of coping programs with all youth provides a basis for lengthier intervention. In particular, Gaylord-Harden and colleagues (2013) found positive effects (.31/–.09) for coping interventions at the conclusion of programming but no or negative effects at follow-up (first follow-up = –.02 /–.09; second follow-up = –.08/–.17). These findings may indicate that lengthier intervention and/or booster sessions are required to help youth generalize or ingrain new coping skills. Additional intervention research is needed to test this hypothesized interpretation. If it proves valid, we may need to shift our typical approach to administering coping interventions from discrete, individually based, brief interventions provided by mental health professionals to intervention that is embedded within ongoing programming of schools or community-based organizations that includes group-based activism focused on systemic problems affecting low-income youth.

Implications of Moderator Data and Related Findings Although results of moderator studies just reviewed provide few consistent messages relevant to intervention, related findings provide evidence that contextual variables can affect the extent to which protective factors are helpful and interventions are effective, particularly within the context of urban poverty. For example, as with the broader coping literature (Compas et al., 2002, 2010), most basic and intervention studies of coping with poverty stressors have reported positive effects for active, problem-solving, engagement approaches and negative effects for avoidance and disengagement (Raviv & Wadsworth, 2010; Wadsworth & Berger, 2006; Wadsworth & Santiago, 2008). Nonetheless, this general pattern is tempered by findings, particularly with youth residing in urban poverty, of positive effects for avoidant, disengagement strategies, at least in the short run (Gaylord-Harden, Taylor, Campbell, Kesselring, & Grant, 2009; Gonzalez et al., 2001). For example, Gaylord-Harden and colleagues (2009) found that youth who behaviorally avoided stressors associated with urban poverty (e.g., exposure to community violence) were better off than those who actively engaged. These findings highlight the importance of matching coping strategy with stressor (Compas et al., 2002, 2010). Future intervention work would benefit from greater integration with basic moderator studies to iteratively determine which methods of coping are most

effective for specific poverty-related stressors. This body of studies also suggests not only that strategies deemed ineffective in other settings may be protective in the context of urban poverty but that generally effective strategies may prove ineffective or even problematic for low-income urban youth (Ceballo, Ramirez, Hearn, & Maltese, 2003; Hammack, Richards, Luo, Edlynn, & Roy, 2004; Li, Nussbaum, & Richards, 2007; Youngstrom, Weist, & Albus, 2003). For example, D'Imperio, Dubow, and Ippolito (2000) reported no protective effects for individual, family, or extrafamilial resources for their sample of low-income urban youth. This pattern fits with Gerard and Buehler's (2004) findings that cumulative risk across multiple domains overwhelms the effects of potential protective factors for youth and with the construct of protective reactive effects, defined by Luthar, Cicchetti, and Becker (2000) as the dissipation of protective effects under conditions of overwhelming stress. More basic and intervention research is needed to determine the contexts in which coping and other protective factors are helpful and harmful. For example, Grant and colleagues found that individually based active coping strategies were, generally, counter-productive for the most severely stressed youth in their low-income urban sample (Grant, McMahon, Carter, & Carleton, in press). However, even the most severely stressed youth could benefit from active strategies if they were used in the context of supportive relationships with adults and connections to protective settings (e.g., family, school, community, or religious organization) (Grant, McMahon, et al., in press). Grant and colleagues are currently building on those findings through the development of a new intervention which teaches low-income urban youth contextually relevant coping strategies and connects them with mentors and protective settings to support their coping efforts (Grant, Farahmand, et al., 2014). This broad approach is consistent with research, reviewed earlier, on the numerous mechanisms through which poverty exerts its negative effects, including reduced interpersonal and community resources. Additional research is needed to evaluate the efficacy and effectiveness of broad approaches such as this one. Meta-analyses of intervention effects also highlight the important moderating function context can play on effectiveness. In particular, Farahmand and colleagues' recent meta-analyses of mental health interventions targeting youth residing in urban poverty (Farahmand, Duffy, et al., 2012; Farahmand, Grant, Polo, Duffy, & Dubois, 2011) indicate that community-based interventions (.25 /–.11) are more effective than school-based interventions (.08 /–.09). School-based interventions in low-income urban communities were particularly likely to yield negligible or even negative effects if they targeted externalizing outcomes. This pattern is consistent with evidence that stressors particularly common within the context of urban poverty (i.e., exposure to community violence) are specifically associated with externalizing outcomes (Grant et al., 2013). In the short run, these findings suggest that community-based interventions should be prioritized over school-based interventions when targeting externalizing problems affecting low-income urban youth. In addition, there is a need for research on systemic barriers to effective schoolbased intervention in the context of urban poverty, particularly for externalizing problems.

Research is needed to determine what school-based factors within the context of urban poverty compromise the delivery of effective mental health intervention and what strategies might be useful in addressing these negative factors so that, in the long run, we can harness the potential benefit of school-based interventions for youth residing in urban poverty (Durlak, Weissberg, Dymnicki, Taylor, & Schellinger, 2011).

Future Directions Earlier we discussed several pathways that appear to be good candidates linking childhood experiences of deprivation with psychological distress and conduct disorders. These pathways included elevated parental stress and subsequent parental harshness and lack of sensitivity as well as children's own stress responses. One of the primary reasons childhood deprivation appears to be negative for children and possibly extending later into the life course is because of exposure to a higher level of cumulative risk factors, such as family turmoil, violence, chaos, and instability as well as suboptimal physical environmental conditions (e.g., substandard housing, toxins) that often accompany poverty. We also pointed out that increasing evidence points toward compromised executive functioning and less adaptive coping strategies as difficulties accompanying childhoods with less opportunity. Finally, we provided evidence that for stress and executive functioning, very recent work has begun to illuminate the role of brain structure (e.g., smaller hippocampus) and function (e.g., more difficulty regulating negative emotions) in relation to childhood poverty/low SES. Thus, not only are low-SES children apt to be exposed to a higher level of aversive environmental demands, they may also be less capable of coping with the myriad of demands they are confronted with at home, in school, and in the community. Low-SES individuals have higher demands but often lack the intrapersonal or environmental resources to manage those demands (Gallo & Matthews, 2003). Given what appears to be the central role of parenting in the linkages between poverty and deprived environments for children's mental health (see Figure 4.1), it is worth thinking about other ways in which poverty and low SES environments could alter the parent-child relationship. We noted that although the focus has been on the direct effects of income or SES on parenting behaviors, some work suggests that deprivation may interact with other critical aspects of parents' lives to moderate the effects of SES on parenting behaviors. In addition to expanding the small amount of work noted earlier on social support, religiosity, and chaos as potential moderators of the link between SES and parenting behavior, another topic worth considering is social capital. Many low-SES families, particularly in urban areas, also reside in neighborhoods low in social capital (Leventhal & Brooks-Gunn, 2000; Saegert, Thompson, & Warren, 2001; Sampson, Morenoff, & Gannon-Rowely, 2002). These settings tend to be low in social cohesion and social control, making it difficult for parents to connect with others who can also contribute to the development and well-being of children. Low-income women have smaller and less cohesive social networks than their middle-income counterparts (Evans, Boxill, & Pinkava, 2008). A related aspect of low social capital is the dearth of positive, informal learning and socialization opportunities, such as music, dance, and art; access to museums and other informal learning centers; after-school activities; summer programs; and

organized sports and recreational activities. Such resources tend to provide greater opportunities for positive and sustained interactions with other adults and increase the probability of constructive, positive peer interactions (Eccles & Gootman, 2002). There is some evidence that high social capital can buffer some of the ill effects of childhood poverty on adolescent smoking and weight gain (Evans & Kutcher, 2011). Recall also that better classroom practices help low-income adolescents better regulate themselves, which, in turn, lead to better psychological adjustment (Brody et al., 2002). At the same time, social capital can indirectly support more positive, productive parenting. It provides opportunities to meet and develop friendships with other adults who may have shared values and aspirations, particularly for their children. A richer network of community resources for children and youth also means their parents get some respite from child care. Evans et al. (2008) showed that the negative relation between family income levels and maternal responsiveness in a sample of low- and middle-income 13-year-olds was largely a by-product of high maternal stress and low social resources available to the mother. Each of these factors was associated with lower maternal responsiveness, but the joint pathway of low social resources and high stress was what really mattered. Furthermore, we suspect many lowSES parents feel uncomfortable and possibly even unwelcome at their child's school. Such perceptions, regardless of their accuracy, when coupled with single parenthood and/or nonstandard working hours that may also be variable, make it much less likely that low-SES parents develop relationships with school personnel or cross paths with other parents at their children's school. Parental engagement in school is strongly tied to SES (Evans, 2004). We need to learn about how community social capital can directly influence low-income children's psychological well-being directly as well as affect their parents' abilities to manage the myriad of demands associated with raising a family. Another contributing risk factor that has not received much attention in the linkages between childhood poverty and the development of psychopathology is stigma and discrimination. Adolescents who are poor feel stigmatized (Fitchen, 1981), which in turn contributes to elevated allostatic load relative to their more advantaged peers (Fuller-Rowell, Evans, & Ong, 2012). This finding is bolstered by the persistent findings that some of the SES-health gradient can be attributed to inequality (Wilkinson & Pickett, 2009) with states or countries having the widest gaps between the rich and the poor also having the steepest SES-health gradients (Kondo et al., 2009). How inequality translates into ill health is a topic ripe for additional research, and it is likely that stigma, bias, and discrimination play a role in SES and child psychopathology. There is also abundant evidence that racial discrimination is stressful and affects the physical and mental health of children, youth, and their parents (Mays, Cochran, & Barnes, 2007). Racial discrimination also explains some of the covariation between between minority status and ill health (Williams, Mohammed, Leavell, & Collins, 2010). The interplay of lower financial opportunity with stigma and discrimination related to poverty itself along with its interaction with racial minority status on children's mental health is an important topic worthy of much more attention. Wilson, Foster, Anderson, and Mance (2009) investigated the role of racial identity but found it did not moderate the link between poverty and psychological distress among African American youth. Grant and colleagues (2005), however, did show that

exposure to violence was a mediator of linkages between poverty and internalizing symptoms among African American adolescents. It is sobering to consider the potential indirect effects of racism on parenting behaviors in the context of poverty. It is hard to be warm and responsive or to build a wide social network when one has to be worried about how one is perceived by others. This situation is likely compounded by simultaneous efforts to be vigilant about how one's children are faring in school or the larger community because of their racial status. The degree of social isolation because of low-SES status and/or racial minority status likely also intersects with urbanity. One of the striking features of urban poverty is the concentration of low-income, racial minority families in specific neighborhoods. This concentration stands in marked contrast to rural areas, where being poor and/or being non-White makes one a distinct demographic minority (Lichter & Johnson, 2007). Another important area for further exploration in research on childhood deprivation and the development of psychopathology is the topic of mastery or control. One of the elements inherent in nearly all of the environmental demands low-SES children and their families face is uncontrollability. A large body of literature documents that chronic exposure to negative social (Peterson, Maier, & Seligman, 1993) and physical (Evans & Stecker, 2004) conditions elevate helplessness in children and adults. Chronic experience of helplessness is a well-documented risk factor for depression (Peterson et al., 1993). Poverty and other forms of disadvantage themselves may lead to elevated experiences of uncontrollability producing helplessness. The lower the family's income, the less children believe they have mastery (Bandura, Barbaranelli, Caprara, & Pastorelli, 2001; Battle & Rotter, 1963). Among adolescents and adults, the higher one's income and education, the higher one's perceived control (Specht, Egloff, & Schmukle, 2013). Moreover, income reductions at any point during a 6-year period led to decreased perceived control, independently of baseline income levels (Specht et al., 2013). Furthermore, lower-SES parents are more authoritarian (Hoff et al., 2002) and less likely to encourage selfdirectedness or autonomy in their children (Kasser, Koestner, & Lekes, 2002; Luster, Rhoades, & Haas, 1989; Whitbeck et al., 1997). Furthermore, in a sample of Head Start children, those faced with a wider array of risk factors in the home environment exhibited more learned helplessness on a standard helplessness protocol (Brown, 2009). In this protocol, participants are given an interesting, age-appropriate puzzle. The index of helplessness is persistence to solve the puzzle. Using a similar protocol, Evans and colleagues (2005) also found that children from ages 9 to13 from low-income households exhibited greater learned helplessness compared to their middle-income peers. Given the prominence of racism and discrimination in the experience of low-income African American children and their families, control is likely an important pathway in thinking about the intersection of race and class as it impacts the mental health of children and youth (Williams et al., 2010). A few investigators have looked at the potential role of genetic processes to influence outcomes of childhood poverty. Early chronic stress exposure is capable of altering genetic transcription processes (i.e., epigenesis). One research team has examined epigenetic processes in relation to childhood SES. Work by Miller, Chen, and Cole (2009) reveals that early low SES childhoods are related to altered gene expression in the inflammatory response

systems (e.g., T cells) as well as in glucocorticoid (i.e., stress hormone) receptor systems (Miller & Chen, 2007; Miller et al., 2009). Heightened reactivity to adverse stimuli is the net effect of these different changes in gene transcription profiles. Recent work by Essex and colleagues (2013) has provided some of the first evidence of differential DNA methylation (affects the protein-encoding portions of a given gene, making it more or less accessible to transcription and expression) in adolescents whose childhoods were marked by high levels of stress and adversity. These epigenetic “marks” on the genome caused by early experience are hypothesized to lead to stable changes in the expression of genes relevant to health and development. We also know that genetic polymorphisms can alter reactivity to stressful environments, and this insight has recently been applied to childhood SES. In a twin study with 5-year-olds, additive, genetic effects accounted for about 70% of the variation in resilience for behavioral adjustment in low-SES, resilient children (Kim-Cohen, Moffitt, Caspi, & Tayor, 2004). Moreover, genetic components also carried some of the contribution of maternal warmth to resilience. One genetic polymorphism that has generated considerable interest in clinical psychology is the serotonin transporter genotype (5-HTTLPR) because it appears to relate directly to psychopathology and can alter vulnerability to risk factor exposure. Recent work suggests some possible linkages between this genetic polymorphism and vulnerability to lowSES background. Juveniles from low-SES backgrounds who had the long/long allele of 5HTTLPR were more prone to narcissism and callous-unemotional traits than youth with the short/short and short/long genotypes, who did not differ in narcissism or callousness as a function of SES (Sadeh et al., 2010). Gene Brody and his colleagues (2002) have been following a sample of low-SES, African American youth from elementary school through young adulthood. Among those at risk, given high levels of poverty-related stressors, a subset of young adults were resilient in terms of good psychological adjustment and academic achievement. One cluster of these resilient young adults, however, manifested very high levels of allostatic load, indicative of chronic physiological stress. The resilient subgroup who did well despite the difficult circumstances in which they grew up appeared to pay a price in the form of very high levels of chronic stress. This physical vulnerability was associated with the short allele genetic polymorphism on the 5-HTTLPR serotonin transporter genotype. Finally, it warrants brief mention that many of the chronic stressors accompanying low-SES childhoods, such as low maternal responsiveness, chaos, and financial pressure, each interact with specific genetic polymorphisms to alter children's vulnerability (Kim & Evans, 2011). The last future research topic we wish to call attention to is the paucity of a developmental focus in much of the research. As we noted in our overviews of research on childhood SES and psychopathology, very few studies incorporate comparisons of the mental health of children of different ages, let alone follow the same child over time in a longitudinal design (see NICHD, 2005; Schoon et al., 2002; Wadsworth & Achenbach, 2005, for some notable exceptions). There are trends in the literature suggesting that both the age and the duration of poverty exposure matter for children's mental health, indicating that earlier poverty is worse than later poverty and that the longer the duration, the more adverse the impacts (Bolger et al., 1995; Brooks-Gunn, Guo, & Furstenberg, 1993; Macmillan, McMorris, & Kruttschnitt, 2004;

Najman et al., 2010; NICHD, 2005; Sacker, Schoon, & Bartley, 2002; Schonberg & Shaw, 2007; Schoon et al., 2002). Furthermore, many epidemiological investigations of SES reveal that social class at birth is a robust predictor of adult morbidity and mortality. These relations hold independently of adulthood SES, and several studies suggest that adult SES has little or no predictive power for health when childhood SES is statistically accounted for (Cohen, Janicki-Deverts, Chen, & Matthews, 2010). Two particularly elegant demonstrations of the developmental timing of poverty and children's psychological well-being come from two experimental interventions. Costello and colleagues (2003, 2010) evaluated the impacts of increased incomes from the opening of a casino on a Native American reservation comparing Native American and non-Native children who did not receive the income supplements from the casino. The supplements reduced externalizing symptoms in the Native American children, but no changes over the same time period occurred in the comparison group. These same differences maintained for the subjects as adults but only for the youngest cohort who were exposed from an earlier age for a longer time period to the income supplements. Fernald and Neufeld (2008) and Fernald and colleagues (2009) evaluated a conditional cash-transfer poverty reduction program conducted in Mexico (Opportunidades). Essentially parents are given cash supplements contingent on various positive child-rearing practices, such as school attendance and use of preventive health care. Low-income rural children and their families were randomly assigned to the interventions at different time periods but were all enrolled at birth. Eight to 10 years after the intervention, children who had been enrolled 6 months earlier in the program showed more significant improvements in psychological health compared to those who had been enrolled later on. The section titled “Practice and Policy” provides more information on conditional cash transfer programs. Overall, however, our knowledge base for understanding the differential effects of of variations in poverty's timing and duration on children's mental health is thin and does not extend over long periods of maturation. Much more work spanning multiple developmental periods and capturing the timing of income changes needs to be done in order to tease apart the differential and/or interactive effects of developmental timing of poverty and duration of exposure to poverty. One reason it is important to examine this question more rigorously is because it speaks to the issue of whether there are particular points in development where the adverse impacts of lower financial and social opportunities matter more and then become embedded in the individual, scarring him or her for life. Or is what really matters the accumulation of multiple insults that recur as one grows up in poverty or in a low-SES household (Shonkoff, Boyce, & McEwen, 2009)? These two models are not mutually exclusive and could both be happening. They do have different implications for interventions. If, for example, there are critical developmental periods during which scarring occurs, interventions occurring prior to these periods may have the largest impacts. If, however, the process of damage occurs over time and continues to build over time, interventions may be helpful at multiple points in development. For example, interventions targeting family income, parental stress, and parenting may be most effective when delivered as early in development as possible, whereas interventions can be added later in the developmental process to directly target older children's self-regulation, EF,

and coping skills and to prevent further accumulation of damage. Developmental issues also intertwine with our presentation of potential, underlying mechanisms to help account for why poverty and low-SES childhoods are so inimical to the development of mental health. For example, if high levels of chronic stress affect different cortical and subcortical areas of the brain at different maturational rates, then we might expect an early critical period to be more consequential for systems that largely mature early in life. In contrast, biological systems that mature over a longer time frame might be more vulnerable to the accumulation of adverse experiences accompanying poverty rather than being age dependent. The PFC and other stress-regulatory regions of the brain, such as the hippocampus, take a longer time to mature than the amgydala (Lupien et al., 2009; Tottenham & Sheridan, 2010). The amygdala is a limbic structure centrally involved in rapid, automated responses to threat, fear, and strong emotions. This way of thinking about poverty, low SES, and child development also has important implications for whether constructs like coping or executive functioning would be expected to mediate the adverse impacts of poverty on mental health or function as moderators, providing protection. Wadsworth, Raviv, Compas, and Connor-Smith (2005) found that coping strategies were more apt to mediate the links between poverty-related stress and mental health among young adolescents but to moderate the same relations for their parents. Mediation would be predicted for systems and processes that mature earlier in life whereas moderation would be predicted for those that come on line later. Consider, for example, the relations between SES and EF functioning in relation to age. As reviewed earlier, low SES tends to directly erode EF among younger children; for adolescents, however, EF often functions as a resource, moderating the ill effects of SES on mental health. A developmental perspective also encourages us to think more about the relative influence of the home environment compared to the larger school and neighborhood environment in the lives of low-SES children. Thus, the impact of poverty directly on parents relative to the impact of poverty on school and community may play out quite differently as a function of child age. Finally, all of our research programs will benefit from better implementation of multiple levels of analysis to bring together converging lines of evidence from the community, neighborhood, family, individual, and genetic realms. An integrated understanding of the risks and sources of resilience in children affected by poverty is likely to lead the best solutions for this social problem that is so very costly, both in terms of human suffering as well as losses of human and economic capital. Our overarching goal in writing this chapter was to create a compendium of the excellent scholarship that has convincingly demonstrated the critical roles of poverty and low socioeconomic status in the development of most forms of psychopathology. What is quite clear from this review is that we have moved beyond the important but insufficient question of Does poverty matter for developmental psychopathology? to the essential question of How does poverty matter for developmental psychopathology? The research outlined here on the psychosocial and biological mechanisms of poverty's effects on a child's brain, body, and functioning is substantial, compelling, and exciting because it illuminates numerous processes and pathways that can be targeted in treatment and prevention. The consistency of findings across multiple levels of analysis and inference provides a strong basis on which to urge

researchers to explicitly include and explore income and SES variables in developmental psychopathology research—controlling for or covarying the effects of income/SES relegates this critical risk factor to a nuisance variable. It is clear that poverty and low SES are more than a nuisance and instead constitute a very powerful determinant of children's mental health.

References Aber, J. L., Bennett, N. G., Conley, D. C., & Li, J. (1997). The effects of poverty on child health and development. Annual Review of Public Health, 18, 463–483. Ackerman, B. P., Brown, E. D., & Izard, C. E. (2004). The relations between contextual risk, earned income, and the school adjustment of children from economically disadvantaged families. Developmental Psychology, 40(2), 204–216. doi: 10.1037/0012–1649.40.2.204 Adler, N. E., Bush, N. R., & Pantell, M. S. (2012). Rigor, vigor, and the study of health disparities. Proceedings of the National Academy of Sciences, 109, 17154–17159. Adler, N. E., & Stewart, J. (2010). Health disparities across the lifespan: Meaning, methods, and mechanisms. Annals of the New York Academy of Sciences, 1186, 5–23. Addy, S., Engelhardt, W., & Skinner, C. (2013). Basic facts about low-income children: Children under 3 years, 2011. New York, NY: National Center on Children in Poverty. Aldwin, C. M. (2007). Stress, coping and development (2nd ed.). New York, NY: Guilford Press. Allensworth, E. (2006). Update to From High School to the Future: A first look at Chicago Public School graduates' college enrollment, college preparation, and graduation from four-year colleges. Chicago, IL: Consortium on Chicago School Research. Almeida, D. M., Neupert, S. D., Banks, S. R., & Serido, J. (2005). Do daily stress processes account for socioeconomic health disparities? Journals of Gerontology: Series B: Psychological Sciences and Social Sciences, 60B(2), 34–39. doi: 10.1093/geronb/60.Special_Issue_2.S34 Amone-P'Olak, K., Burger, H., Ormel, J., Huisman, M., Verhulst, F. C., & Oldehinkel, A. J. (2009). Socioeconomic position and mental health problems in pre- and early-adolescents: The TRAILS study. Social Psychiatry and Psychiatric Epidemiology, 44(3), 231–238. doi: 10.1007/s00127-008-0424-z Ardila, A., Rosselli, M., Matute, E., & Guajardo, S. (2005). The influence of parents' educational level on the development of executive functions. Developmental Neuropsychology, 28, 539–560. Attar, B., Guerra, N., & Tolan, P. (1994). Neighborhood disadvantage, stressful life events and adjustments in urban elementary school children. Journal of Child Clinical Psychology, 23,

391–400. Bandura, A., Barbaranelli, C., Caprara, G. V., & Pastorelli, C. (2001). Self-efficacy beliefs as shapers of children's aspirations and career trajectories. Child Development, 72, 187–206. Barker, E. D., Copeland, W., Maughan, B., Jaffee, S. R., & Uher, R. (2012). Relative impact of maternal depression and associated risk factors on offspring psychopathology. The British Journal of Psychiatry, 200(2), 124–129. doi: 10.1192/bjp.bp.111.092346 Battle, E. S., & Rotter, J. B. (1963). Children's feelings of personal control as related to social class and ethnic group. Journal of Personality, 31, 482–490. Belle, D. (1990). Poverty and women's mental health. American Psychologist, 45, 385–389. Bendersky, M., & Lewis, M. (1994). Environmental risk, biological risk, and developmental outcome. Developmental Psychology, 30, 484–494. Beidas, R. S., Suarez, L., Simpson, D., Read, K., Wei, C., Connolly, S., & Kendall, P. (2012). Contextual factors and anxiety in minority and European American youth presenting for treatment across two urban university clinics. Journal of Anxiety Disorders, 26(4), 544–554. Bierman, K. (2002). Evaluation of the first 3 years of the Fast Track prevention trial with children at high risk for adolescent conduct problems. Journal of Abnormal Child Psychology, 30(1), 19–35. Blair, C. (2010). Stress and the development of self-regulation in context. Child Development Perspectives, 3, 181–188. Blair, C., Granger, D. A., Willoughby, M., Mills-Koonce, R., Cox, M., Greenberg, M. T., et al. (2011). Salivary cortisol mediates effects of poverty and parenting on executive functions in early childhood. Child Development, 82, 1970–1984. Blair, C., & Raver, C. C. (2012). Child development in the context of adversity. American Psychologist, 67, 309–318. Blair, C., Raver, C., Granger, D., Mills-Koonce, R., & Hibel, L. (2011). Allostasis and allostatic load in the context of poverty in early childhood. Development and Psychopathology, 23(3), 845–857. doi: 10.1017/S0954579411000344 Bøe, T., Øverland, S., Lundervold, A. J., & Hysing, M. (2012). Socioeconomic status and children's mental health: Results from the bergen child study. Social Psychiatry and Psychiatric Epidemiology, 47(10), 1557–66. doi: 10.1007/s00127–011–0462–9 Bolger, K. E., Patterson, C., Thompson, W., & Kupersmidt, J. (1995). Psychosocial adjustment among children experiencing persistent and intermittent family economic hardship. Child Development, 66, 1107–1129. Bond, M. A., & Keys, C. B. (1993). Empowerment, diversity, and collaboration: Promoting

synergy on community boards. American Journal of Community Psychology, 21, 37–57. Bornstein, M. H. (1989). Sensitive periods in development: Structural characteristics and causal interpretations. Psychological Bulletin, 105(2), 179–197. doi: 10.1037/00332909.105.2.179 Bornstein, M. H. (Ed.). (1989). Maternal responsiveness: Characteristics and consequences. San Francisco, CA: Jossey-Bass. Bradley, R. H. (2003). Environment and parenting. In M. H. Bornstein (Ed.), Handbook of parenting, Volume 2. Biology and ecology of parenting (Vol. 2, pp. 281–314). Mahwah, NJ: Erlbaum. Bradley, R. H., & Corwyn, R. F. (2002). Socioeconomic status and child development. Annual Review of Psychology, 53(1), 371–399. doi: 10.1146/annurev.psych.53.100901.135233 Bradley, R. H., Corwyn, R. F., Burchinal, M., McAdoo, H. P., & Garcia-Coll, C. (2001). The home environment of children in the United States Part II: Relations with behavioral development through age thirteen. Child Development, 72, 1868–1886. Bradley, R. H., Corwyn, R. F., McAdoo, H. P., & Garcia-Coll, C. (2001). The home environments of children in the United States Part I: Variations by age, ethnicity, and poverty status. Child Development, 72, 1844–1867. Braveman, P., Egerter, S., & Williams, D. R. (2011). The social determinants of health: Coming of age. Annual Review of Public Health, 32, 381–398. Brody, G. H., Chen, Y. F., Kogan, S. M., Yu, T., Molgaard, V. K., DiClemente, R. J., & Wingood, G. M. (2012). Family-centered program deters substance use, conduct problems, and depressive symptoms in black adolescents. Pediatrics, 129(1), 108–115. Brody, G. H., Dorsey, S., Forehand, R., & Armistead, L. (2002). Unique and protective contributions of parenting and classroom processes to the adjustment of African American children living in single-parent families. Child Development, 73, 274–286. Brody, G. H., & Flor, D. L. (1997). Maternal psychological functioning, family processes, and child adjustment in rural, single-parent, African American families. Developmental Psychology, 33, 1000–1011. Brody, G. H., Murry, V. M., Gerrard, M., et al. (2006). The Strong African American Families program: Prevention of youths' high-risk behavior and a test of a model of change. Journal of Family Psychology, 20(1), 1–11. Brody, G. H., Stoneman, Z., & Flor, D. (1996). Parental religiosity, family processes, and youth competence in rural, two-parent African American families. Developmental Psychology, 32, 696–706. Brody, G. H., Yu, T., Chen, Y. F., Kogan, S. M., Evans, G. W., Beach, S. R. H., et al. (2013).

Cumulative socioeconomic status risk, allostatic load, and adjustment: A prospective latent profile analysis with contextual and genetic protective factors. Developmental Psychology, 49, 913–927. Brooks-Gunn, J. (2003). Do you believe in magic?: What we can expect from early childhood intervention programs. Social Policy Report, 17(1), 1–16. Brooks-Gunn, J., & Duncan, G. J. (1997). The effects of poverty on children. Future of Children, 7(2), 55–71. Retrieved from http://search.proquest.com/docview/619359309? accountid=13158 Brooks-Gunn, J., Guo, G., & Furstenberg, F. F., Jr. (1993). Who drops out of and who continues beyond high school? A 20-year followup of black urban youth. Journal of Research on Adolescence, 3, 271–294. Brooks-Gunn, J., Klebanov, P., & Liaw. (1995). The learning, physical, and emotional environment of the home in the context of poverty: The infant health and development program. Children Youth and Services Review, 17, 251–276. Brotman, L. M., Calzada, E., Huang, K. Y., Kingston, S., Dawson-McClure, S., Kamboukos, D.,… & Petkova, E. (2011). Promoting effective parenting practices and preventing child behavior problems in school among ethnically diverse families from underserved, urban communities. Child Development, 82(1), 258–276. Brown, E. D. (2009). Persistence in the fact of academic challenge for economically disadvantaged children. Early Childhood Research, 7, 175–186. Brown, L., Cowen, E., Hightower, A., & Lotyczewski, B. (1986). Demographic differences among children in judging and experiencing specific life events. Journal of Special Education, 20, 339–346. Buckner, J. C., Mezzacappa, E., & Beardslee, W. R. (2003). Characteristics of resilient youths living in poverty: The role of self-regulatory processes. Development and Psychopathology, 15, 139–162. Buckner, J. C., Mezzacappa, E., & Beardslee, W. R. (2009). Self-regulation and its relations to adaptive functioning in low income youths. American Journal of Orthopsychiatry, 79, 19–30. Burchinal, M. R., Roberts, J. E., Hooper, S., & Zeisel, S. A. (2000). Cumulative risk and early cognitive development: A comparison of statistical risk models. Developmental Psychology, 36, 793–807. Butterworth, P., Cherbuin, N., Sachdev, P., & Anstey, K. J. (2012). The association between financial hardship and amygdala and hippocampal volumes: Results from the PATH through life project. Social Cognitive and Affective Neuroscience, 7, 548–556. Ceballo, R., Ramirez, C., Hearn, K. D., & Maltese, K. L. (2003). Community violence and

children's psychological well-being: Does parental monitoring matter? Journal of Clinical Child and Adolescent Psychology, 32, 586–592. Chen, E., Hanson, M. D., Paterson, L. Q., Griffin, M. J., Walker, H. A., & Miller, G. E. (2006). Socioeconomic status and inflammatory processes in childhood asthma: the role of psychological stress. Journal of Allergy and Clinical Immunology, 117(5), 1014–1020. Chen, E., Langer, D. A., Raphaelson, Y. E., & Matthews, K. A. (2004). Socioeconomic status and health in adolescents: The role of stress interpretations. Child Development, 75, 1039– 1052. Chen, E., & Matthews, K. A. (2001). Cognitive appraisal biases: An approach to understanding the relation between socioeconomic status and cardiovascular reactivity in children. Annals of Behavioral Medicine, 23, 101–111. Chen, E., Matthews, K. A., & Boyce, W. T. (2002). Socioeconomic differences in children's health: How and why do these relationships change with age? Psychological Bulletin, 128, 295–329. Chen, E., & Miller, G. E. (2012). Shift and persist strategies: Why low socioeconomic status isn't always bad for health. Perspectives on Psychological Science, 7, 135–158. Chen, E., Miller, G. E., Kobor, M. S., & Cole, S. W. (2011). Maternal warmth buffers the effects of low early-life socioeconomic status on pro-inflammatory signaling in adulthood. Molecular Psychiatry, 16, 729–737. Chen, E., Miller, G. E., Lachman, M. E., Gruenewald, T. L., & Seeman, T. E. (2012). Protective factors for adults from low-childhood socioeconomic circumstances: The benefits of shift-and-persist for allostatic load. Psychosomatic Medicine, 74, 178–186. Chen, E., Miller, G. E., Walker, H. A., Arevalo, J. M., Sung, C. Y., & Cole, S. W. (2009). Genome-wide transcriptional profiling linked to social class in asthma. Thorax, 64, 38–43. Chen, E., Strunk, R. C., Trethewey, B. A., Schreier, H. M. C., Maharaj, N., & Miller, G. E. (2011). Resilience in low-socioeconomic-status children with asthma: Adaptations to stress. Journal of Allergy and Clinical Immunology, 128, 970–976. Chicago Public Schools. (n.d.). Retrieved July 17, 2013, from http://www.cps.edu/SchoolData/Pages/SchoolData.aspx Chilton, M., Chyatte, M., & Breaux, J. (2007). The negative effects of poverty and food insecurity on child development. Indian Journal of Medical Research, 126, 262–272. Church, W. T., Jaggers, J. W., & Taylor, J. K. (2012). Neighborhood, poverty, and negative behavior: An examination of differential association and social control theory. Children and Youth Services Review, 34(5), 1035–1041. doi: 10.1016/j.childyouth.2012.02.005 Cleveland, H. (2003). Disadvantaged neighborhoods and adolescent aggression: Behavioral

genetic evidence of contextual effects. Journal of Research on Adolescence, 13(2). Cohen, J. S., Edmunds, J. M., Brodman, D. M., Benjamin, C. L., & Kendall, P. C. (2012). Using self-monitoring: Implementation of collaborative empiricism in cognitive-behavioral therapy. Cognitive and Behavioral Practice. Cohen, S., Janicki-Deverts, D., Chen, E., & Matthews, K. A. (2010). Childhood socioeconomic status and adult health. Annals of the New York Academy of Sciences, 1186, 37–55. Coie, J. D., Dodge, K. A., & Damon, W. (1998). Aggression and antisocial behavior. Hoboken, NJ: Wiley. Retrieved from http://search.proquest.com/docview/620712503? accountid=13158 Coldwell, J., Pike, A., & Dunn, J. (2006). Household chaos—links with parenting and child behaviour. Journal of Child Psychology and Psychiatry, 47, 116–1122. Coleman-Jensen, A., Nord, M., Andrews, M., and Carlson, S. (2011, September). Household Food Security in the United States in 2010. ERR-125, U.S. Department of Agriculture, Economic Research Service. Compas, B. E., Champion, J. E., Forehand, R., Cole, D. A., Reeslund, K. L., Fear, J.,… Roberts, L. (2010). Coping and parenting: Mediators of 12-month outcomes of a family group cognitive–behavioral preventive intervention with families of depressed parents. Journal of Consulting and Clinical Psychology, 78, 623–634. Compas, B. E., Connor-Smith, J. K., Saltzman, H., Thomsen, A. H., & M. E. Wadsworth, M. E. (2002). Coping with stress during childhood and adolescence: Problems, progress, and potential in theory and research. Psychological Bulletin, 127, 87–127. Conger, R. D., Conger, K. J., Elder, G. H., Jr., Lorenz, F, Simons, R., & Whitbeck, L. (1992). A family process model of economic hardship and adjustment of early adolescent boys. Child Development, 63, 526–541 Conger, R. D., Conger, K. J., & Martin, M. J. (2010). Socioeconomic status, family processes, and individual development. Journal of Marriage and Family, 72(3), 685–704. Conger, R. D., & Donnellan, M. B. (2007). An interactionist perspective on the socioeconomic context of human development. Annual Review of Psychology, 58, 175–199. Conger, R. D., & Elder, G. H., Jr. (1994). Families in troubled times. New York, NY: Aldine de Gruyter. Conger, R. D., Elder, G. H., Jr., Lorenz, E., Conger, K., Simons, R., Whitbeck, L.,… Melby, J. (1990). Linking economic hardship to marital quality and instability. Journal of Marriage and the Family, 52, 643–656. Cook, J., & Jeng, K. (2009). Child food insecurity: The economic impact on our Nation.

Chicago, IL: Feeding America. Cooper, J. (2007). Cognitive dissonance: 50 years of a classic theory. London, UK: Sage. Costello, E. J., Compton, S. N., Keeler, G., & Angold, A. (2003). Relationship between poverty and psychopathology: A natural experiment. Journal of the American Medical Association, 290, 2023–2029. Costello, E. J., Erkanli, A., Copeland, W., & Angold, A. (2010). Association of family income supplements in adolescence with development of psychiatric and substance use disorders in adulthood among an American Indian population. Journal of the American Medical Association, 303, 1954–1960. Costello, E. J., Keeler, G. P., & Arnold, A. (2001). Poverty, race/ethnicity, and psychiatric disorder: A study of rural children. American Journal of Public Health, 91, 1494–1498. Cote, S., Vaillancourt, T., LeBlanc, J. C., Nagin, D. S., & Tremblay, R. E. (2006). The development of physical aggression from toddlerhood to pre-adolescence: A nation-wide longitudinal study of Canadian children. Journal of Abnormal Child Psychology, 34(1), 68– 82. Crook, S. R., & Evans, G. W. (2014). The role of planning skills in the income–achievement gap. Child Development, 85(2), 405–411. Cushon, J. A., Vu, L. H., Janzen, B. L., & Muhajarine, N. (2011). Neighborhood poverty impacts children's physical health and well-being over time: Evidence from the Early Development Instrument. Early Education and Development, 22(2), 183–205. doi: 10.1080/10409280902915861 Dahl, Gordon B., and Lance Lochner. 2012. “The Impact of Family Income on Child Achievement: Evidence from the Earned Income Tax Credit.” American Economic Review, 102(5): 1927–56. doi: 10.1257/aer.102.5.1927 Dalton, J. H., Hill, J., Thomas, E., & Kloos, B. (2013). Community psychology. In D. K. Freedheim & I. B. Weiner (Eds.), Handbook of psychology, Vol. 1: History of psychology (2nd ed., pp. 468–487). Hoboken, NJ: Wiley. Darley, J. M., & Latané, B. (1968). Bystander intervention in emergencies: diffusion of responsibility. Journal of Personality and Social Psychology, 8, 377–383 Dearing, E., McCartney, K., & Taylor, B. A. (2006). Within-child associations between family income and externalizing and internalizing problems. Developmental Psychology, 42(2), 237– 252. doi: 10.1037/0012–1649.42.2.237 Deater-Deckard, K., Chen, N. C., Wang, Z., & Bell, M. A. (2012). Socioeconomic risk moderates the link between household chaos and maternal executive function. Journal of Family Psychology, 26, 391–399.

Deater-Deckard, K., Dodge, K. A., Bates, J. E., & Pettit, G. S. (1998). Multiple risk factors in the development of externalizing behavior problems: Group and individual differences. Development and Psychopathology, 10, 469–493. DeNavas-Walt, C., Proctor, B. D., &. Smith, J. C. (2012). Income, Poverty, and Health Insurance Coverage in the United States: 2011. U.S. Census Bureau, Current Population Reports, P60–243, Washington, DC: U.S. Government Printing Office. Deng, S., Lopez, V., Roosa, M. W., Ryu, E., Burrell, G. L., Tein, J. Y., & Crowder, S. (2006). Family processes mediating the relationship of neighborhood disadvantage to early adolescent internalizing symptoms. Journal of Early Adolescence, 26, 206–231. Dickens, W. T., Sawhill, I., & Tebbs, J. (2006). The effects of investing in early childhood education on economic growth. Brookings Working Paper). Washington, DC: The Brookings Institution. Diemer, M. A., Mistry, R. S., Wadsworth, M. E., López, I., & Reimers, F. (2013). Best practices in conceptualizing and measuring social class in psychological research. Analyses of Social Issues and Public Policy, 13(1), 77–113. D'Imperio, R. L., Dubow, E. F., & Ippolito, M. F. (2000). Resilient and stress-affected adolescents in an urban setting. Journal of Clinical Child Psychology, 29, 129–142. Doan, S. N., Fuller-Rowell, T. E., & Evans, G. W. (2012). Cumulative risk and adolescent's internalizing and externalizing problems: The mediating roles of maternal responsiveness and self-regulation. Developmental Psychology, 48, 1529–1539. Dodge, K. A., & Conduct Problems Prevention Research Group. (2007). Fast track randomized controlled trial to prevent externalizing psychiatric disorders: Findings from grades 3 to 9. Journal of the American Academy of Child & Adolescent Psychiatry, 46(10), 1250–1262. Dohrenwend, B. P., Levav, I., Shrout, P. E., Schwartz, S., Naveh, G., Link, B. G.,… Stueve, A. (1992). Socioeconomic status and psychiatric disorders: The causation-selection issue. Science, 255(5047), 946–952. Retrieved from http://search.proquest.com/docview/618135117?accountid=13158 D'Onofrio, B. M., Goodnight, J. A., Hulle, C. A., Rodgers, J. L., Rathouz, P. J., Waldman, I. D., & Lahey, B. B. (2009). A quasi-experimental analysis of the association between family income and offspring conduct problems. Journal of Abnormal Child Psychology, 37(3), 415– 429. Downey, G., & Coyne, J. (1990). Children of depressed parents. Psychological Bulletin, 108, 50–76. DuBois, D. L., Felner, R. D., Meares, H., & Krier, M. (1994). Prospective investigation of socioeconomic disadvantage, life stress, and social support on early adolescent adjustment.

Journal of Abnormal Psychology, 103, 511–522. Dubow, E., Tisak, J., Casey, D., Hryshko, A., & Reid, G. (1991). A two-year longitudinal study of stressful life events, social support, and social problem solving skills: Contributions to children's behavioral and academic adjustment. Child Development, 62, 583–599. Duncan, G. J., Brooks-Gunn, J., & Klebanov, P. (1994). Economic deprivation and early childhood development. Child Development, 65(2), 296–318. Duncan, G. J. & Petersen, E. (2001). The long and short of asking questions about income, wealth and labor supply. Social Science Research, 30, 248–263. Duncan, G. J., Ziol-Guest, K., & Kalil, A. (2010). Early-childhood poverty and adult attainment, behavior, and health. Child Development, 81(1), 306–325. doi: 10.1111/j.1467– 8624.2009.01396.x Durlak, J. A., Weissberg, R. P., Dymnicki, A. B., Taylor, R. D., & Schellinger, K. B. (2011). The impact of enhancing students' social and emotional learning: A meta-analysis of schoolbased universal interventions. Child Development, 82, 405–432. Eamon, M. (2000). Structural model of the effects of poverty on externalizing and internalizing behaviors of four- to five-year-old children. Social Work Research, 24(3), 143–154. Eccles, J. S., & Gootman, J. A. (Eds.). (2002). Community programs to promote youth development. Washington, DC: National Academies Press. Eckenrode, J., Campa, M., Luckey, D. W., Henderson, C. R., Cole, R., Kitzman, H.,… Olds, D. (2010). Long-term effects of prenatal and infancy nurse home visitation on the life course of youths: 19-Year follow-up of a randomized trial. Archives of Pediatric and Adolescent Medicine, 164(1), 9. doi: 10.1001/archpediatrics.2009.240 Ekberg-Aronsson, M., Nilsson, P. M., Nilsson, J. A., Löfdahl, C. J. & Löfdahl, K. (2007). Mortality risks among heavy-smokers with special reference to women: A long-term follow-up of an urban population. European Journal of Epidemiology, 22(5), 301–309. Elder, G. H., Jr. (1974). Children of the great depression. Chicago, IL: University of Chicago Press. Elder, G. H., Jr., vanNguyen, T., & Caspi, A. (1985). Linking family hardship to children's lives. Child Development, 56, 361–375. Essex, M. J., Boyce, W. T., Hertzman, C., Lam, L. L., Armstrong, J. M., Neumann, S. M. A., & Kobor, M. S. (2013). Epigenetic vestiges of early developmental adversity: Childhood stress exposure and DNA methylation in adolescence. Child Development, 84(1), 58–75. doi: 10.1111/j.1467-8624.2011.01641.x Evans, G. W. (2004). The environment of childhood poverty. American Psychologist, 59, 77– 92.

Evans, G. W., Boxhill, L., & Pinkava, M. (2008). Poverty and maternal responsiveness: The role of maternal stress and social resources. International Journal of Behavioral Development, 32, 232–237. Evans, G. W., Chen, E., Miller, G. E., & Seeman, T. E. (2012). How poverty gets under the skin: A life course perspective. In V. Maholmes & R. King (Eds.), The Oxford handbook of poverty and child development (pp. 13–36). New York, NY: Oxford University Press. Evans, G. W., Eckenrode, J., & Marcynyszyn, L. A. (2010). Poverty and chaos. In G. W. Evans & T. D. Wachs (Eds.), Chaos and its influence on children's development: An ecological perspective (pp. 225–238). Washington, DC: American Psychological Association. Evans, G. W., & English, K. (2002). The environment of poverty: Multiple stressor exposure, psychophysiological stress, and socioemotional adjustment. Child Development, 73, 1238– 1248. Evans, G. W., Gonnella, C., Marcynyszyn, L. A., Gentile, L., & Salpekar, N. (2005). The role of chaos in poverty and children's socioemotional adjustment. Psychological Science, 16, 560–565. Evans, G. W., & Kim, P. (2007). Childhood poverty and health: Cumulative risk exposure and stress dysregulation. Psychological Science, 18, 953–957. Evans, G. W., & Kim, P. (2010). Multiple risk exposure as a potential explanatory mechanism for the socioeconomic status-health gradient. Annals of the New York Academy of Sciences, 1186, 174–189. Evans, G. W., & Kim, P. (2012). Childhood poverty and young adult allostatic load: The Mediating role of childhood cumulative risk exposure. Psychological Science, 23, 979–983. Evans, G. W., Kim, P. K., Ting, A. H., Tesser, H. B., & Shanis, D. (2007). Cumulative risk, maternal responsiveness, and allostatic load among young adolescents. Developmental Psychology, 43, 341–351. Evans, G. W., & Kutcher, R. (2011). Loosening the links between childhood poverty and adolescent smoking and obesity: The protective effects of social capital. Psychological Science, 22, 3–7. Evans, G. W., Li, D., & Whipple, S. S. (2013). Cumulative risk and child development. Psychological Bulletin, 139, 1342–1396. Evans, G. W., & Rosenbaum, J. (2008). Self-regulation and the income-achievement gap. Early Childhood Research Quarterly, 23, 504–514. Evans, G. W., & Schamberg, M. A. (2009). Childhood poverty, chronic stress, and adult working memory. Proceedings of the National Academy of Sciences, 106, 6545–6549.

Evans, G. W., & Stecker, R. (2004). The motivational consequences of environmental stress. Journal of Environmental Psychology, 24, 143–165. Evans, W., Wolfe, B., & Adler, N. (2012). The SES and health gradient: A brief review of the literature. Wolfe, B., Evans, W. and Seeman, T. The Biological Consequences of Socioeconomic Inequalities. New York: Russell Sage Foundation. Fanti, K. A., & Henrich, C. C. (2010). Trajectories of pure and co-occurring internalizing and externalizing problems from age 2 to age 12: Findings from the national institute of child health and human development study of early child care. Developmental Psychology, 46(5), 1159– 1175. doi: 10.1037/a0020659 Farah, M. J., Shera, D. M., Savage, J. H., Betancourt, L., Giannetta, J. M., Brodsky, N. L.,… & Hurt, H. (2006). Childhood poverty: Specific associations with neurocognitive development. Brain research, 1110(1), 166–174. Farahmand, F. K., Duffy, S. N., Tailor, M. A., DuBois, D. L., Lyon, A. L., Grant, K. E.,… & Nathanson, A. M. (2012). Community-Based Mental Health and Behavioral Programs for Low-Income Urban Youth: A Meta-Analytic Review. Clinical Psychology: Science and Practice, 19(2), 195–215. Farahmand, F. K., Grant, K. E., Polo, A., Duffy, S. N., & Dubois, D. L. (2011). School-based mental health programs for low-income urban youth: A systematic and meta-analytic review. Clinical Psychology: Science and Practice, 18, 372–390. Federman, M., Garner, T. I., Short, K., Cutter, W. B. I., Kiely, J., Levine, D., et al. (1996, May). What does it mean to be poor in America? Monthly Labor Review, 3–17. Felner, R. D., Brand, S., DuBois, D. L., Adan, A., Mulhall, P., & Evans, E. (1995). Socioeconomic disadvantage, proximal environmental experiences and socioemotional and academic adjustment in early adolescence: An investigation of mediated effects model. Child Development, 66, 774–792. Fernald, L. C., Gertler, P. J., & Neufeld, L. M. (2008). Role of cash in conditional cash transfer programmes for child health, growth, and development: an analysis of Mexico's Oportunidades. The Lancet, 371(9615), 828–837. Fernald, L. C. H., Gertler, P. J., & Neufeld, L. M. (2009). 10-year effect of Oportunidades, Mexico's conditional cash transfer programme, on child growth, cognition, language, and behavior: A longitudinal follow-up study. Lancet, 374, 1997–2005. Festinger, L. (1957/1985). A theory of cognitive dissonance. Stanford, CA: Stanford University Press. Fielding, K., & Head, B. (2012). Determinants of young Australians' environmental actions: The role of responsibility attributions, locus of control, knowledge and attitudes. Environmental Education Research, 18(2),171–186.

Fiese, B. H. (2006). Family routines and rituals. New Haven, CT: Yale University Press. Fiese, B. H., & Winter, M. A. (2010). The dynamics of family chaos and its relation to children's socio-emotional well being. In G. W. Evans & T. D. Wachs (Eds.), Chaos and children's development: Levels of analysis and mechanisms (pp. 49–66). Washington, DC: American Psychological Association. Fine, S. (2002, September). The wealth of neighborhoods: The effects of a housing mobility project on perceptions of neighborhood safety, maternal depression, parenting behavior and youth externalizing behavior. Dissertation Abstracts International Section A, 63. Fitchen, J. M. (1981). Poverty in rural America. Prospect Heights, IL: Waveland. Fiszbein, A., Schady, N. R., & Ferreira, F. H. (2009). Conditional cash transfers: reducing present and future poverty. World Bank Publications. Frank, D. A., Casey, P. H., Black, M. M., Rose-Jacobs, R., Chilton, M., Cutts, D., &… Cook, J. T. (2010). Cumulative hardship and wellness of low-income, young children: Multisite surveillance study. Pediatrics, 125(5), e1115–e1123. doi: 10.1542/peds.2009–1078 Fuller-Rowell, T. E., Evans, G. W., & Ong, A. D. (2012). Poverty and health: The mediating role of perceived discrimination. Psychological Science, 23, 734–739. Furnham, A. (2003). Belief in a just world: Research progress over the past decade. Personality and Individual Differences, 34, 795–817. Galindo, C., & Fuller, B. (2010). The social competence of Latino kindergartners and growth in mathematical understanding. Developmental Psychology, 46(3), 579–592. doi: 10.1037/a0017821 Gallo, L. C., & Matthews, K. A. (2003). Understanding the association between socioeconomic status and physical health: Do negative emotions play a role? Psychological Bulletin, 129, 10–31. Ganzel, B. L., Morris, P. A., & Wethington, E. (2010). Allostasis and the human brain: Integrating models of stress from the social and life sciences. Psychological Review, 117, 134–174. Garner, P. W., & Spears, F. M. (2000). Emotion regulation in low-income preschoolers. Social Development, 9, 246–264. Gaylord-Harden, N., Duffy, S., Doxie, J., Nolan, L., DuBois, D., Grant, K. E.,… Tolan, P. H. (2013). The effects of coping interventions on psychosocial functioning in youth: A metaanalytic review. Manuscript under review. Gaylord-Harden, N. K., Taylor, J. J., Campbell, C. L., Kesselring, C. M., & Grant, K. E. (2009). Maternal attachment and depressive symptoms in urban adolescents: The influence of coping strategies and gender. Journal of Clinical Child and Adolescent Psychology, 38, 684–

695. Gaylord-Harden, N. K., Gipson, P. M., Mance, G., & Grant, K. E. (2008). Coping patterns of low-income urban African American youth: A confirmatory factor analysis and cluster analysis of The Children's Coping Strategies Checklist. Psychological Assessment, 20, 10–22. Gennetian, L. A., & Miller, C. (2002). Children and welfare reform: A view from an experimental welfare program in Minnesota. Child Development, 73, 601–620. Georgiades, K., Boyle, M. H., & Duku, E. (2007). Contextual influences on children's mental health and school performance: The moderating effects of family immigrant status. Child Development, 78, 1572–1591. Gerard, J. M., & Buehler, C. (2004). Cumulative environmental risk and youth problem behavior. Journal of Marriage and Family, 66, 702–720. Gershoff, E. T., Aber, J. L., Raver, C. C., & Lennon, M. C. (2007). Income is not enough: Incorporating material hardship into models of income associations with parenting and child development. Child Development, 78, 70–95. Gianaros, P. J., Horenstein, J. A., Cohen, S., Matthews, K. A., Brown, S. M., Flory, J. D.,… & Hariri, A. R. (2007). Perigenual anterior cingulate morphology covaries with perceived social standing. Social Cognitive and Affective Neuroscience, 2(3), 161–173. Gianaros, P. J., Horenstein, J. A., Hariri, A. R., Sheu, L. K., Manuck, S. B., Matthews, K. A., & Cohen, S. (2008). Potential neural embedding of parental social standing. Social cognitive and affective neuroscience, 3(2), 91–96. Gianaros, P. J., & Manuck, S. B. (2010). Neurobiological pathways linking socioeconomic position and health. Psychosomatic Medicine, 72, 450–461. Gianaros, P. J., Manuck, S. B., Sheu, L. K., Kuan, D. C., Votruba-Drzal, E., Craig, A. E., & Hariri, A. R. (2010). Parental education predicts corticostriatal functionality in adulthood. Cerebral Cortex, bhq160. Godfrey, B., & Lowe, C. (1975). Devaluation of innocent victims: An attribution analysis within the just world paradigm. Journal of Personality and Social Psychology, 31, 944–951. Gonzales, N. A., Tein, J. Y., Sandler, I. N., & Friedman, R. J. (2001). On the limits of coping: Interaction between stress and coping for inner-city adolescents. Journal of Adolescent Research, 16, 372–395. Goodnight, J. A., Lahey, B. B., Van Hulle, C. A., Rodgers, J. L., Rathouz, P. J., Waldman, I. D., & D'Onofrio, B. M. (2012). A quasi-experimental analysis of the influence of neighborhood disadvantage on child and adolescent conduct problems. Journal of Abnormal Psychology, 121(1), 95–108. doi: 10.1037/a0025078 Grant, K. E., Compas, B. E., Stuhlmacher, A. F., Thurm, A. E., McMahon, S. D., & Halpert, J.

A. (2003). Stressors and child and adolescent psychopathology: Moving from markers to mechanisms of risk. Psychological Bulletin, 129, 447–466. Grant, K. E. Compas, B. E., Thurm, A. E., McMahon, S. Gipson, P. Campbell, A.,… Westerholm, R. I. (2006). Stressors and child and adolescent psychopathology: Evidence of moderating and mediating effects. Clinical Psychology Review, 26, 257–283. Grant, K. E., Farahmand, F., Meyerson, D. A., Dubois, D. L., Tolan, P. H., Gaylord-Harden, N. K.,… & Duffy, S. (2014). Development of Cities Mentor Project: An Intervention to Improve Academic Outcomes for Low-Income Urban Youth Through Instruction in Effective Coping Supported by Mentoring Relationships and Protective Settings. Journal of prevention & intervention in the community, 42(3), 221–242. Grant, K. E., McMahon, S. D., Carter, J. S., & Carleton, R. (in press). The influence of stressors on the development of psychopathology. Handbook of developmental psychopathology. Grant, K. E., McCormick, A., Poindexter, L., Simpkins, T., Janda, C. M., Thomas, K. J.,… & Taylor, J. (2005). Exposure to violence and parenting as mediators between poverty and psychological symptoms in urban African American adolescents. Journal of Adolescence, 28, 507–521. Grant, K. E., O'Koon, J. H., Davis, T. H., Roache, N. A., Poindexter, L. M., Armstrong, M. L., … & McIntosh, J. M. (2000). Protective factors affecting low-income urban African American youth exposed to stress. Journal of Early Adolescence, 20, 388–417. Grant, K. E., Wagstaff, A., Thomas, K. J., Carlson, G. C., Gipson, P. Y., Mance, G. A.,… Lambert, S. F. (2013). Specific processes linking economic stressors, exposure to community violence, and psychological symptoms affecting low-income urban youth. Manuscript submitted for publication. Green, J. G., McLaughlin, K. A., Berglund, P. A., Gruber, M. J., Sampson, N. A., Zaslavsky, A. M.,… & Kessler, R. C. (2010). Childhood adversities and adult psychiatric disorders in the National Comorbidity Survey Replication I. Archives of General Psychiatry, 67, 113–123. Greenberg, M. T., Lengua, L. J., Coie, J. D., & Pinderhughes, E. E. (1999). Predicting developmental outcomes at school entry using a multiple-risk model: Four American communities. Developmental Psychology, 35, 403–417. Gruenewald, T. L., Cohen, S., Matthews, K. A., Tracy, R., & Seeman, T. E. (2009). Association of socioeconomic status with inflammation markers in Black and White men and women in the Coronary Artery Risk Development in Young Adults (CARDIA) study. Social Science & Medicine, 69(3), 451–459. doi: 10.1016/j.socscimed.2009.05.018 Hackman, D. A., & Farah, M. J. (2009). Socioeconomic status and the developing brain. Trends in Cognitive Science, 13, 65–73.

Hackman, D. A., Farah, M. J., & Meaney, M. J. (2010). Socioeconomic status and the brain: Mechanistic insights from human and animal research. Nature Neuroscience, 11, 651–659. Hamlin, J., Wynn, K., Bloom, P., & Mahajan, N. (2011). How infants and toddlers react to antisocial others. Proceedings of the National Academy of Sciences of the United States of America, 108(50), 19931–19936. Hammack, P. L., Richards, M. H., Luo, Z., Edlynn, E. S., & Roy, K. (2004). Social support factors as moderators of community violence exposure among inner-city African American young adolescents. Journal of Clinical Child and Adolescent Psychology, 34, 450–462. Hammack, P. L., Robinson, W. L. V., Crawford, I., & Li, S. T. (2004). Poverty and depressed mood among urban African-American adolescents: A family stress perspective. Journal of Child and Family Studies, 13, 309–323. Hanson, J. L., Chandra, A., Wolfe, B. L., & Pollak, S. D. (2011). Association between income and the hippocampus. PLoS One, 6, 1–8. Hanson, M. J., Miller, A. D., Diamond, K., Odom, S., Lieber, J., Butera, G., &… Fleming, K. (2011). Neighborhood community risk influences on preschool children's development and school readiness. Infants & Young Children, 24(1), 87–100. doi: 10.1097/IYC.0b013e3182008dd0. Hao, L., & Matsueda, R. L. (2006). Family dynamics through childhood: A sibling model of behavior problems. Social Science Research, 35(2), 500–524. doi: 10.1016/j.ssresearch.2004.10.003 Harnish, J. D., Dodge, K. A., & Valente, E. (1995). Mother-child interaction quality as a partial mediator of the roles of maternal depressive symptomatology and socioeconomic status in the development of child behavior problems. Child Development, 66(3), 739–753. Retrieved from http://search.proquest.com/docview/618682429?accountid=13158 Hart, W., Albarracín, D., Eagly, A. H., Brechan, I., Lindberg, M. J., & Merrill, L. (2009). Feeling validated versus being correct: A meta-analysis of selective exposure to information. Psychological Bulletin, 135(4), 555–588. Hashima, P. Y., & Amato, P. (1994). Poverty, social support, and prenatal behavior. Child Development, 65, 394–403. Haskins, R., & Rouse, C. E. (2013). Time for change: A new federal strategy to prepare disadvantaged students for college. Future of Children, 2, 1–5. Haynie, D. L., Silver, E., & Teasdale, B. (2006). Neighborhood characteristics, peer networks, and adolescent violence. Journal of Quantitative Criminology, 22(2), 147–169. doi: 10.1007/s10940–006–9006-y Heatherton, T. F. (2011). Neuroscience of self and self-regulation. Annual Review of

Psychology, 62, 363–390. Heckman, J. J. (2006). Skill formation and economics of investing in disadvantaged children. Science, 312, 1900–1902. Henninger, W. R., & Luze, G. (2013). Moderating effects of gender on the relationship between poverty and children's externalizing behavior. Journal of Child Health Care, 17, 72–81. Hill, A. L., Degnan, K. A., Calkins, S. D., & Keane, S. P. (2006). Profiles of externalizing behavior problems for boys and girls across preschool: The roles of emotion regulation and inattention. Developmental Psychology, 42(5), 913–928. doi: 10.1037/0012–1649.42.5.913 Hillemeier, M. M., Lanza, S. T., Landale, N. S., & Oropesa, R. S. (2012). Measuring early childhood health and health disparities: A new approach. Maternal and Child Health Journal. doi: 10.1007/s10995–012–1205–6 Hinshaw, S. P., Lahey, B. B., & Hart, E. L. (1993). Issues of taxonomy and comorbidity in the development of conduct disorder. Development and Psychopathology, 5(1), 31. Retrieved from http://search.proquest.com/docview/1292827461?accountid=13158 Hofer, C., Eisenberg, N., & Reiser, M. (2010). The role of socialization, effortful control, and ego resiliency in french adolescents' social functioning. Journal of Research on Adolescence, 20(3), 555–582. doi: 10.1111/j.1532-7795.2010.00650.x Hoff, E., Laursen, B., & Tardif, T. (2002). Socioeconomic status and parenting. In M. H. Bornstein (Ed.), Handbook of parenting (2nd ed., pp. 231–252). Mahwah, NJ: Erlbaum. Holzer, H. J., Schanzenbach, D., Duncan, G. J., & Ludwig, J. (2008). The economic costs of childhood poverty in the United States. Journal of Children & Poverty, 14(1) 41–61. doi: 10.1080/10796120701871280 Howse, R. B., Lange, G., Farran, D. C., & Boyles, C. D. (2003). Motivation and selfregulation as predictors of achievement in economically disadvantaged children. Journal of Experimental Education, 71, 151–174. Hughes, C., Ensor, R., Wilson, A., & Graham, A. (2010). Tracking executive function across the transition to school: A latent variable approach. Developmental Neuropsychology, 35, 20– 36. Iceland, J. (2003). Why poverty remains high: The role of income growth, economic inequality, and changes in family structure, 1949–1999. Demography, 40(3), 499–519. Isaacs, J., & Magnuson, K. (2011). Income and education as predictors of children's school readiness. Washington, DC: Brookings Institution. Janoff-Bulman, R., Timko, C., & Carli, L. L. (1985). Cognitive biases in blaming the victim. Journal of Experimental Social Psychology, 21(2), 161–177.

Jorm, A. (2012). Mental health literacy: Empowering the community to take action for better mental health. American Psychologist, 67(3), 231–243. Jyoti, D. F., Frongillo, E. A., & Jones, S. J. (2005). Food insecurity affects school children's academic performance, weight gain, and social skills. Journal of Nutrition, 135, 2831–2839. Juster, R. P., Bizik, G., Picard, M., Arsenault-Lapierre, G., Sindi, S., Trepanier, L.,… & Lupien, S. J. (2011). A transdisciplinary perspective on chronic stress in relation to psychopathology throughout life span development. Development and Psychopathology, 23, 725–776. Kahn, J. R., & Pearlin, L. I. (2006). Financial strain over the life course and health among older adults. Journal of Health and Social Behavior, 47(1), 17–31. doi: 10.1177/002214650604700102 Kainz, K., Willoughby, M. T., Vernon-Feagans, L., & Burchinal, M. R. (2012). Modeling family economic conditions and young children's development in rural United States: Implications for poverty research. Journal of Family and Economic Issues, 33(4), 410–420. doi: 10.1007/s10834–012–9287–2 Karoly, Lynn A., M. Rebecca Kilburn and Jill S. Cannon. Proven Benefits of Early Childhood Interventions. Santa Monica, CA: RAND Corporation, 2005. http://www.rand.org/pubs/research_briefs/RB9145. Karriker-Jaffe, K., Foshee, V. A., Ennett, S. T., & Suchindran, C. (2009). Sex differences in the effects of neighborhood socioeconomic disadvantage and social organization on rural adolescents' aggression trajectories. American Journal of Community Psychology, 43(3–4), 189–203. doi: 10.1007/s10464–009–9236-x Kasser, T., Koestner, R., & Lekes, N. (2002). Early family experiences and adult values: A 26year, prospective longitudinal study. Personality and Social Psychology Bulletin, 28, 826– 835. Keller, T. E., Spieker, S. J., & Gilchrist, L. (2005). Patterns of risk and trajectories of preschool problem behaviors: A person-oriented analysis of attachment in context. Development and Psychopathology, 17(2), 349–384. doi: 10.1017/S0954579405050170 Kendler, K. S., & Eaves, L. J. (1986). Models for the joint effect of genotype and environment on liability to psychiatric illness. American Journal of Psychiatry, 143(3), 279–289. Retrieved from http://search.proquest.com/docview/617147622?accountid=13158 Kim, P., & Evans, G. W. (2011). Family resources, genes, and human development. In A. Booth, S. M. Mc Hall. & N. S. Lansdale (Eds.), Biosocial foundations of family processes (pp. 221–230). New York, NY: Springer. Kim, P., Evans, G. W., Angstadt, M., Ho, S. S., Sripada, C. S., Swain, J. E.,… & Phan, K. L. (2013). Effects of childhood poverty and chronic stress on emotion regulatory brain function in

adulthood. Proceedings of the National Academy of Sciences, 110(46), 18442–18447. Kim-Cohen, J., Moffitt, T. E., Caspi, A., & Taylor, A. (2004). Genetic and environmental processes in young children's resilience and vulnerability to socioeconomic deprivation. Child Development, 75, 651–668. Kingston, S., Huang, K. Y., Calzada, E., Dawson-McClure, S., & Brotman, L. (2013). Parent involvement in education as a moderator of family and neighborhood socioeconomic context on school readiness among young children. Journal of Community Psychology, 41(3), 265– 276. Kishiyama, M. M., Boyce, W. T., Jimenez, A. M., Perry, L. E., & Knight, R. T. (2008). Socioeconomic disparities affect prefrontal function in children. Journal of Cognitive Neuroscience, 21, 1106–1115. Kjellstrand, J. M., & Eddy, J. M. (2011). Parental incarceration during childhood, family context, and youth problem behavior across adolescence. Journal of Offender Rehabilitation, 50(1), 18–36. doi: 10.1080/10509674.2011.536720 Knudsen, E. I., Heckman, J., Cameron, J., & Shonkoff, J. P (2006). Economic, neurobiological, and behavioral perspectives on building America's future workforce. Proceedings of the National Academy of Sciences, 103, 10155–10162. Kohen, D. E., Leventhal, T., Dahinten, V., & McIntosh, C. N. (2008). Neighborhood disadvantage: Pathways of effects for young children. Child Development, 79(1), 156–169. doi: 10.1111/j.1467–8624.2007.01117.x Kondo, N., Sembajwe, G., Kawachi, I., van Dam, R. M., Subramanian, S. V., & Yamagata, Z. (2009). Income inequality, mortality, and self rated health: meta-analysis of multilevel studies. Bmj, 339, b4471. Kowaleski-Jones, L., & Duncan, G. J. (1999). The structure of achievement and behavior across middle childhood. Child Development, 70(4), 930–943. Retrieved from http://search.proquest.com/docview/619404234?accountid=13158 Krishnakumar, A., & Black, M. M. (2002). Longitudinal predictors of competence among African American children: The role of distal and proximal risk factors. Journal of Applied Developmental Psychology, 23(3), 237–266. Kroneman, L., Loeber, R., & Hipwell, A. E. (2004). Is neighborhood context differently related to externalizing problems and delinquency for girls compared with boys? Clinical Child and Family Psychology Review, 7(2), 109–122. doi: 10.1023/B:CCFP.0000030288.01347.a2 Lansford, J. E., Malone, P. S., Stevens, K. I., Dodge, K. A., Bates, J. E., & Pettit, G. S. (2006). Developmental trajectories of externalizing and internalizing behaviors: Factors underlying resilience in physically abused children. Development and Psychopathology, 18(1), 35–55.

Retrieved from http://search.proquest.com/docview/201697661?accountid=13158 Lantz, P. M., Lynch, J. W., House, J. S., Lepkowski, J. M., Mero, R. P., Musick, M. A., & Williams, D. R. (2001). Socioeconomic disparities in health change in a longitudinal study of US adults: The role of health-risk behaviors. Social Science and Medicine, 53, 29–40. Latané, B., & Darley, J. M. (1968). Group inhibition of bystander intervention in emergencies. Journal of Personality and Social Psychology, 10, 215–221. Leary, M. R., & Forsyth, D. R. (1987). Attributions of responsibility for collective endeavors. Review of Personality and Social Psychology, 8, 167–188. Lee, B. J., & Mackey-Bilaver, L. (2007). Effects of WIC and food stamp program participation on child outcomes. Children and Youth Services Review, 29(4), 501–517. doi: 10.1016/j.childyouth.2006.10.005 Lengua, L. J. (2012). Poverty, the development of effortful control, and children's academic, social, and emotional adjustment. In V. Maholmes & R. B. King (Eds.), The Oxford handbook of poverty and child development (pp. 491–511). New York, NY: Oxford University Press. Lengua, L. J., Bush, N. R., Long, A. C., Kovacs, E. A., & Trancik, A. M. (2008). Effortful control as a moderator of the relation between contextual risk factors and growth in adjustment problems. Development and Psychopathology, 20, 509–528. Lerner (1980). The belief in a just world: A fundamental delusion. New York, NY Plenum Press. Leventhal, T., & Brooks-Gunn, J. (2000). The neighborhoods they live in: the effects of neighborhood residence on child and adolescent outcomes. Psychological Bulletin, 126, 309– 337. Leventhal, T., & Brooks-Gunn, J. (2011). Changes in neighborhood poverty from 1990 to 2000 and youth's problem behaviors. Developmental Psychology, 47(6), 1680–1698. Leventhal, T. M., & Newman, S. (2010). Housing and child development. Children and Youth Services Review, 32, 1165–1174. Li, S. T., Nussbaum, K. M., & Richards, M. H. (2007). Risk and protective factors for urban African American youth. American Journal of Community Psychology, 39, 21–35. Liaw, F., & Brooks-Gunn, J. (1994). Cumulative familial risks and low birth weight children's cognitive and behavioral development. Journal of Clinical Child Psychology, 23, 360–372. Lichter, D. T., & Johnson, K. M. (2007). The changing spatial concentration of America's rural poor population. Rural Sociology, 72, 331–358. Li-Grining, C. P. (2007). Effortful control among low-income preschoolers in three cities: Stability, change, and individual differences. Developmental Psychology, 43, 208–221.

Lipina, S. J., Martelli, M. I., Vuelta, B., & Colombo, J. A. (2005). Performance on the A-not-B task of Argentina infants from unsatisfied and satisfied basic needs. Interamerican Journal of Psychology, 39, 49–60. Lloyd, J. E., Li, L., & Hertzman, C. (2010). Early experiences matter: lasting effect of concentrated disadvantage on children's language and cognitive outcomes. Health Place,16(2), 371–380. Lodewijkx, H. M., de Kwaadsteniet, E. W., & Nijstad, B. A. (2005). That could be me (or not): Senseless violence and the role of deservingness, victim ethnicity, person identification, and position identification. Journal of Applied Social Psychology, 35(7), 1361–1383. Loseman, A., & van den Bos, K. (2012). A self-regulation hypothesis of coping with an unjust world: Ego-depletion and self-affirmation as underlying aspects of blaming of innocent victims. Social Justice Research, 25(1), 1–13. Lupien, S. J., McEwen, B. S., Gunnar, M., & Heim, C. (2009). Effects of stress throughout the lifespan on the brain, behavior, and cognition. Nature Reviews Neuroscience, 10, 434–445. Luster, T., Rhoades, K., & Haas, B. (1989). The relation between parental values and parenting behavior: A test of the Kohn hypothesis. Journal of Marriage and the Family, 51, 139–147. Luthar, S. S. (1999). Poverty and children's adjustment. Thousand Oaks, CA: Sage. Luthar, S. S. (2003). The culture of affluence: Psychological costs of material wealth. Child Development, 74(6), 1581–1593. doi: 10.1046/j.1467–8624.2003.00625.x Luthar, S. S., Cicchetti, D., & Becker, B. (2000). The construct of resilience: A critical evaluation and guidelines for future work. Child Development, 71, 543–562. Lynch, J. W., Kaplan, G. A., & Shema, S. J. (1997). Cumulative impact of sustained economic hardship on physical, cognitive, psychological, and social functioning. New England Journal of Medicine, 337(26), 1889–1895. Retrieved from http://search.proquest.com/docview/619173979?accountid=13158 Macmillan, R., McMorris, B. J., & Kruttschnitt, C. (2004). Linked lives: Stability and change in maternal circumstances and trajectories of antisocial behavior in children. Child Development, 75, 205–220. Magnuson, K., & Duncan, G. J. (2002a). Off with Hollingshead. In M. H. Bornstein & R. H. Bradley (Eds.), Socioeconomic status, parenting, and child development, pp. 83–106. Mahwah, NJ: Erlbaum. Magnuson, K. A., & Duncan, G. J. (2002b). Parents in poverty. In M. H. Bornstein (Ed.), Handbook of parenting (Vol. 4, pp. 95–121). Mahwah, NJ: Erlbaum. Månsson, N. O., Råstam, L., Eriksson, K. F., & Israelsson, B. (1998). Socioeconomic

inequalities and disability pension in middle-aged men. International Journal of Epidemiology, 27(6), 1019–1025. Manuck, S. B., Bleil, M. E., Petersen, K. L., Flory, J. D., Mann, J. J., Ferrell, R. E.,… & Muldoon, M. F. (2005). The socio-economic status of communities predicts variation in brain serotonergric responsivity. Psychosomatic Medicine, 35, 519–528. Manuck, S. B., Flory, J. D., Ferrell, R. E., & Muldoon, M. F. (2004). Socio-economic status covaries with central nervous system serotonergic responsivity as a function of allelic variation in the serotonin transporter gene-linked polymorphic region. Psychoneuroendocrinology, 29, 651–668. Masters, R. D., & Gruter, M. (1992). The sense of justice: Biological foundations of law. Thousand Oaks, CA: Sage. Matthews, K. A., Flory, J. D., Muldoon, M., & Manuck, S. B. (2000). Does socioeconomic status relate to central sertononergic responsivity in healthy adults? Psychosomatic Medicine, 62, 231–237. Matthews, K. A., & Gallo, L. C. (2010). Psychological perspectives on pathways linking socioeconomic status and physical health. Annual Review of Psychology, 61, 501–530. Mayer, S., & Jencks, C. (1989). Poverty and the distribution of material hardship. Journal of Human Resources, 24, 88–114. Mays, V. M., Cochran, S. D., & Barnes, N. W. (2007). Race, race-based discrimination, and health outcomes among African Americans. Annual Review of Psychology, 58, 201–225. McBride Murry, V., Berkel, C., Gaylord-Harden, N. K., Copeland-Linder, N., & Nation, M. (2011). Neighborhood poverty and adolescent development. Journal of Research on Adolescence, 21(1), 114–128. McEwen, B. S. (1998). Protective and damaging effects of stress mediators. New England Journal of Medicine, 338, 171–179. McEwen, B. S. (2012). Brain on stress: How the social environment gets under the skin. Proceedings of the National Academy of Sciences, 109, 17180–17185. McEwen, B. S., & Gianaros, P. J. (2010). Central role of the brain in stress and adaptation: Links to socioeconomic status, health, and disease. Annals of the New York Academy of Sciences, 1186, 190–222. McLaughlin, K. A., Green, J. G., Alegría, M., Costello, E. J., Gruber, M. J., Sampson, N. A., & Kessler, R. C. (2012). Food insecurity and mental disorders in a national sample of U.S. adolescents. Journal of the American Academy of Child & Adolescent Psychiatry, 51(12), 1293–1303. Retrieved from http://search.proquest.com/docview/1269431548? accountid=13158

McLeod, J. D., & Shanahan, M. J. (1993). Poverty, parenting, and children's mental health. American Sociological Review, 58(3), 351–366. McLeod, J. D., & Shanahan, M. J. (1996). Trajectories of poverty and children's mental health. Journal of Health and Social Behavior, 37(3), 207–220. McLoyd, V. C. (1997). Children in poverty: Development, public policy, and practice. In W. Damon, I. Sigel, & A. Renninger (Eds.), Handbook of child psychology. Vol. 4: Child psychology in practice (pp. 135–208). New York, NY: Wiley. McLoyd, V. C. (1998). Socioeconomic disadvantage and child development. American Psychologist, 53, 185–204. Merikangas, K. R., He, J. P., Burstein, M., Swendsen, J., Avenevoli, S., Case, B.,… & Olfson, M. (2011). Service utilization for lifetime mental disorders in US adolescents: results of the National Comorbidity Survey–Adolescent Supplement (NCS-A). Journal of the American Academy of Child & Adolescent Psychiatry, 50(1), 32–45. Mezzacappa, E. (2004). Alerting, orienting, and executive intelligence: Developmental properties and sociodemographic correlates in an epidemiological sample of young, urban children. Child Development, 75, 1373–1386. Miech, R. A., Caspi, A., Moffitt, T. E., Wright, B. R. E., & Silva, P. A. (1999). Low socioeconomic status and mental disorders: A longitudinal study of selection and causation during young adulthood. American Journal of Sociology, 104(4), 1096–1131. Retrieved from http://search.proquest.com/docview/619431908?accountid=13158 Miller, G. E., & Chen, E. (2007). Unfavorable socioeconomic conditions in early life presage expression of proinflammatory phenotype in adolescence. Psychosomatic Medicine, 69, 402– 409. Miller, G., Chen, E., & Cole, S. W. (2009). Health psychology: Developing biologically plausible models linking the social world and physical health. Annual Review of Psychology, 60, 501–524. doi: 10.1146/annurev.psych.60.110707.163551 Miller, G. E., Chen, E., Fok, A. K., Walker, H., Lim, A., Nicholls, S.,… & Kobor, M. S. (2009). Low early-life social class leaves a biological residue manifested by decreased glucocorticoid and increased proinflammatory signaling. Proceedings of the National Academy of Sciences, 106, 14716–14721. Miller, G. E., Lachman, M. E., Chen, E., Gruenewald, T. L., Karlamangla, A. S., & Seeman, T. E. (2011). Pathways to resilience: Maternal nurturance as a buffer against the effects of childhood poverty on metabolic syndrome at midlife. Psychological Science, 22, 1591–1599. Moffitt, T. E. (2005). The new look of behavioral genetics in developmental psychopathology: gene-environment interplay in antisocial behaviors. Psychological bulletin, 131(4), 533.

Moilanen, K. L., Shaw, D. S., Dishion, T. J., Gardner, F., & Wilson, M. (2009). Predictors of longitudinal growth in inhibitory control in early childhood. Social Development, 19, 326– 347. Montada, L., & Lerner, M. (Eds.) (1998) Responses to victimizations and belief in the just world. New York, NY: Plenum Press. Morris, P. A., Duncan, G. J., & Clark-Kauffman, E. (2005). Child well being in an era of welfare reform: The sensitivity of transitions in development to policy change. Developmental Psychology, 41, 919–932. Morris, P. A., & Gennetian, L. A. (2003). Identifying the effects of income on children's development using experimental data. Journal of Marriage and Family, 65(3), 716–729. Mrug, S., & Windle, M. (2009). Mediators of neighborhood influences on externalizing behavior in preadolescent children. Journal of Abnormal Child Psychology, 37, 265–280. Murray, J. D., Spadafore, J., & McIntosh, W. D. (2005). Belief in a just world and social perception: Evidence for automatic activation. Journal of Social Psychology, 145(1), 35–47. Muscatell, K. A., Morelli, S. A., Falk, E. B., Way, B. M., Pfeifer, J. H., Galinsky, A. D.,… & Eisenberger, N. I. (2012). Social status modulates neural activity in the mentalizing network. Neuroimage, 60, 1771–1777. Mynatt, C., & Sherman, S. J. (1975). Responsibility attribution in groups and individuals: A direct test of the diffusion of responsibility hypothesis. Journal of Personality and Social Psychology, 32, 1111–1118. Najman, J. M., Clavarino, A., McGee, T. R., Bor, W., Williams, G. M., & Hayatbakhsh, M. R. (2010). Timing and chronicity of family poverty and development of unhealthy behaviors in children: A longitudinal study. Journal of Adolescent Health, 46(6), 538–544. doi: 10.1016/j.jadohealth.2009.12.001 Najman, J. M., Clavarino, A., McGee, T. R., Bor, W., Williams, G. M., & Hayatbakhsh, M. R. (2010). Timing and chronicity of family poverty and development of unhealthy behaviors in children: A longitudinal study. Journal of Adolescent Health, 46(6), 538–544. doi: 10.1016/j.jadohealth.2009.12.001 Najman, J. M., Hayatbakhsh, M. R., Clavarino, A., Bor, W., O'Callaghan, M. J., & Williams, G. M. (2010). Family poverty over the early life course and recurrent adolescent and young adult anxiety and depression: A longitudinal study. American Journal of Public Health, 100(9), 1719–1723. Neuendorf, C., Kim, P., & Evans, G. W. (2009). Childhood poverty and the development of coping styles. Paper presented at the Society for Research in Child Development. April, 2009. National Institute of Child Health and Human Development. (2005). Duration and

developmental timing of poverty and children's cognitive and social development from birth through third grade. [NICHD Early Child Care Research Network]. Child Development, 76, 795–810. Noble, K. G., Houston, S. M., Kan, E., & Sowell, E. R. (2012). Neural correlates of socioeconomic status in the developing human brain. Developmental Science, 15, 516–527. Noble, K. G., McCandiliss, B. D., & Farah, M. J. (2007). Socioeconomic gradients predict individual differences in neurocognitive abilities. Developmental Science, 10, 464–480. Noble, K. G., Norman, M. F., & Farah, M. J. (2005). Neurocognitive correlates of socioeconomic status in kindergarten children. Developmental Science, 8, 74–87. Nuru-Jeter, A. M., Sarsour, K., Jutte, D. P., & Boyce, W. T. (2010). Socioeconomic predictors of health and development in middle childhood: Variations by socioeconomic status measure and race. Issues in Comprhensive Pediatric Nursing, 33, 59–81. Obradovic, J., Shaffer, A., & Masten, A. S. (2012). Risk and adversity in developmental psychopathology: Progress and future directions. In L. C. Mayes & M. Lewis (Eds.), The environment of human development: A handbook of theory and measurement. (pp. 35–57). New York, NY: Cambridge University Press. Odgers, C. L., Caspi, A., Russell, M. A., Sampson, R. J., Arseneault, L., & Moffit, T. E. (2012). Supportive parenting mediates neighborhood socioeconomic disparities in children's antisocial behavior from ages 5–12. Development and Psychopathology, 24(3), 705–721. doi: 10.1017/S0954579412000326 Olds, D. L., Sadler, L., & Kitzman, H. (2007). Programs for parents of infants and toddlers: Recent evidence from randomized trials. Journal of Child Psychology and Psychiatry, 48(3– 4), 355–391. doi: 10.1111/j.1469–7610.2006.01702.x Otero, G. A. (1994). EEG spectral analysis in children with sociocultural handicaps. International Journal of Neuroscience, 79, 213–220. Otero, G. A. (1997). Poverty, cultural disadvantage and brain development: A study of preschool children in Mexico. Electroencephalography and Clinical Neurophysiology, 102, 512–516. Overholser, J. C. (2011). Collaborative empiricism, guided discovery, and the Socratic method: Core processes for effective cognitive therapy. Clinical Psychology: Science and Practice, 18(1), 62–66. Pachter, L. M., Auinger, P., Palmer, R., & Weitzman, M. (2006). Do parenting and the home environment, maternal depression, neighborhood, and chronic poverty affect child behavioral problems differently in different racial-ethnic groups? Pediatrics, 117(4), 1329–1338. Palmer, S., & Williams, H. (2013). Cognitive behavioral approaches. In J. Passmore, D. B.

Peterson, & T. Freire (Eds.) The Wiley-Blackwell handbook of the psychology of coaching and mentoring (pp. 319–338). Hoboken, NJ: Wiley-Blackwell. Park, J., Fertig, A. R., & Allison, P. D. (2011). Physical and mental health, cognitive development, and health care use by housing status of low-income young children in 20 American cities: A prospective cohort study. American Journal of Public Health, 101(Suppl. 1), S255–S261. Parker, G. R., Cowen, E. L., Work, W. C., & Wyman, P. A. (1990). Test correlates of stress resilience among urban school children. Journal of Primary Prevention, 11, 19–35. Paschall, M. J., & Hubbard, M. L. (1998). Effects of neighborhood and family stressors on African American male adolescents' self-worth and propensity for violent behavior. Journal of Consulting and Clinical Psychology, 66(5), 825–831. doi: 10.1037/0022–006X.66.5.825 Peterson, C., Maier, R., & Seligman, M. E. P. (1993). Learned helplessness. New York, NY: Oxford University Press. Petterson, S. M., & Albers, A. B. (2001). Effects of poverty and maternal depression on early child development. Child Development, 72, 1794–1813. Pungello, E. P., Kainz, K., Burchinal, M., Wasik, B. H., Sparling, J. J., Ramey, C. T., & Campbell, F. A. (2010). Early educational intervention, early cumulative risk, and the early home environment as predictors of young adult outcomes within a high-risk sample. Child Development, 81, 410–426. doi: 10.1111/j.1467–8624.2009.01403.x Pungello, E. P., Kupersmidt, J. B., Burchinal, M. R., & Patterson, C. J. (1996). Environmental risk factors and children's achievement from middle childhood to early adolescence. Developmental Psychology, 32, 755–767. Raizada, R. D. S., & Kishiyama, M. M. (2010). Effects of socioeconomic status on brain development, and how cognitive neuroscience may contribute to leveling the playing field. Frontiers of Neuroscience, 4, 1–11. Ramey, C. T., & Ramey, S. L. (1998). Prevention of intellectual disabilities: early interventions to improve cognitive development. Preventive medicine, 27(2), 224–232. Raver, C. C. (2004). Placing emotional self-regulation in sociocultural and socioeconomic contexts. Child Development, 75, 346–353. Raver, C. C., Blair, C., & Willoughby, M. 2013). Poverty as a predictor of 4-year-olds' executive function: New perspectives on models of differential susceptibility. Developmental Psychology, 49, 292–304. Raver, C. C., Gershoff, E. T., & Aber, J. L. (2007). Testing equivalence of mediating model of income, parenting, and school readiness for White, Black, and Hispanic children in a national sample. Child Development, 78, 96–115.

Raviv, T., & Wadsworth, M. E. (2010). The efficacy of a pilot prevention program for children and caregivers coping with economic strain. Cognitive Therapy and Research, 34(3), 216– 228. Rehkopf, D. H., & Buka, S. L. (2006). The association between suicide and the socioeconomic characteristics of geographical areas: A systematic review. Psychological Medicine, 36(2), 145–157. doi: 10.1017/S003329170500588X Reising, M. M., Watson, K. H., Hardcastle, E. J., Merchant, M. J., Roberts, L., Forehand, R., & Compas, B. E. (2013). Parental depression and economic disadvantage: The role of parenting in associations with internalizing and externalizing symptoms in children and adolescents. Journal of Child and Family Studies, 22(3), 335–343. doi: 10.1007/s10826–012–9582–4 Repetti, R. L., Taylor, S. E., & Seeman, T. E. (2002). Risky families: Family social environments and the mental and physical health of offspring. Psychological Bulletin, 128, 330–366. Riccio, J., Dechausay, N., Greenberg, D., Miller, C., Rucks, Z., & Verma, N. (2010, March). Toward reduced poverty across generations: Early findings from New York City's conditional cash transfer program. New York, NY: MDRC. Retrieved from http://www.mdrc.org/sites/default/files/full_588.pdf Rinsky, J. R., & Hinshaw, S. P. (2011). Linkages between childhood executive functioning and adolescent social functioning and psychopathology in girls with ADHD. Child Neuropsychology, 17(4), 368–390. doi: 10.1080/09297049.2010.544649 Ripple, C. H., & Zigler, E. (2003). Research, policy, and the federal role in prevention initiatives for children. American Psychologist, 58(6–7), 482. Rivera, J., Sotres-Alvarez, D., Habicht, J. P., Shamah, T., & Villalpando, S. (2004). Impact of the Mexican Program for Education, Health, and Nutrition (Progresa) on rates of growth and anemia in infants and young children: A randomized effectiveness study. Journal of the American Medical Association, 291, 2563–2570. Roosa, M. W., Deng, S., Ryu, E., Lockhart Burrell, G., Tein, J. Y. Jones, S.,… & Crowder, S. (2005). Family and child characteristics linking neighborhood context and child externalizing behavior. Journal of Marriage and Family, 67, 515–529. Rutter, M. (1979). Protective factors in children's responses to stress and disadvantage. In M. W. Kent & J. E. Rolf (Eds.), Primary prevention of psychopathology (Vol. 3, pp. 49–74). Hanover, NH: University Press of New England. Sacker, A., Schoon, I., & Bartley, M. (2002). Social inequality in educational achievement and psychosocial adjustment throughout childhood: Magnitude and mechanisms. Social Science and Medicine, 55, 863–880. Sadeh, N., Javdani, S., Jackson, J. J., Reynolds, E. K., Potenza, M. N., Gelernter, J.,… &

Verona, E.(2010). Serotonin transporter gene associations with psychopathic traits in youth vary as a function of socioeconomic status. Journal of Abnormal Psychology, 119, 604–609. Saegert, S., Thompson, J. P., & Warren, M. R. (Eds.). (2001). Social capital and poor communities. New York, NY: Russell Sage Foundation. Sameroff, A. J. (2006). Identifying risk and protective factors for healthy development. In A. Clarke-Stewart & J. Dunn (Eds.), Families count (pp. 53–76). New York, NY: Cambridge University Press. Sameroff, A. J., Bartko, W. T., Baldwin, A., Baldwin, C., & Seifer, R. (1998). Family and social influences on the development of child competence. In M. Lewis & C. Feiring (Eds.), Families, risk, and competence (pp. 161–186). Mahwah, NJ: Erlbaum. Sampson, R. J., Morenoff, J. D., & Gannon-Rowley, T. (2002). Assessing neighborhood effects: Social processes and new directions in research. Annual Review of Sociology, 28, 443–478. Santiago, C. D., Wadsworth, M. E., & Stump, J. (2011). Socioeconomic status, neighborhood disadvantage, and poverty-related stress: Prospective effects on psychological symptoms among diverse low-income families. Journal of Economic Psychology, 32, 218–230. Sapolsky, R. M. (2004). Social status and health in humans and other animals. Annual Review of Anthropology, 33, 393–418. doi: 10.1146/annurev.anthro.33.070203.144000 Sarsour, K., Sheridan, M., Jutte, D., Nuru-Jeter, A., Hinshaw, S., & Boyce, W. T. (2011). Family socioeconomic status and child executive functions: The roles of language, home environment, and single parenthood. Journal of the International Neuropsychological Society, 17, 120–132. Schonberg, M. A., & Shaw, D. S. (2007). Risk factors for boys' conduct problems in poor and lower-middle-class neighborhoods. Journal of Abnormal Child Psychology, 35(5), 759–72. doi: 10.1007/s10802–007–9125–4 Schoon, I., Sacker, A., & Bartley, M. (2003). Socioeconomic adversity and psychological adjustment: A developmental-contextual perspective. Social Science and Medicine, 57, 1001– 1015. Sektnan, M., McClelland, M. M., Acock, A. C., & Morrison, F. J. (2010). Relations between early family risk, children's behavioral regulation, and academic achievement. Early Childhood Research Quarterly, 25, 464–479. Sheridan, M. A., Sarsour, K., Jutte, D., D'Esposito, M., & Boyce, W. T. (2012). The impact of social disparity on prefrontal function in childhood. PLoS One, 7, 1–13. Shonkoff, J. P., Boyce, W. T., & Mc Ewen, B. S. (2009). Neuroscience, molecular biology, and the childhood roots of health disparities. Journal of the American Medical Association, 301,

2252–2259. Shonkoff, J. P., & Garner, A. S. (2012). The lifelong effects of early childhood adversity and toxic stress. Pediatrics, 129, e232–e246. Shonkoff, J. P., & Phillips, D. A. (Eds.). (2000). From neurons to neighborhoods: The science of early childhood development. Washington, DC: National Academies Press. Slopen, N., Fitzmaurice, G., Williams, D. R., & Gilman, S. E. (2010). Poverty, food insecurity, and the behavior for childhood internalizing and externalizing disorders. Journal of the American Academy of Child & Adolescent Psychiatry, 49(5), 444–452. Specht, J., Egloff, B., & Schmukle, S. C. (2013). Everything under control? The effects of age, gender, and education on trajectories of perceived control in a nationally representative German sample. Developmental Psychology, 49, 353–364. Speer, P. W., Peterson, N., Armstead, T. L., & Allen, C. T. (2013). The influence of participation, gender and organizational sense of community on psychological empowerment: The moderating effects of income. American Journal of Community Psychology, 51(1–2), 103–113. Spence, S. H., Najman, J. M., Bor, W., O'Callaghan, M. J., & Williams, G. M. (2002). Maternal anxiety and depression, poverty and marital relationship factors during early childhood as predictors of anxiety and depressive symptoms in adolescence. Journal of Child Psychology and Psychiatry, 43, 457–469. Staff, R. T., Murray, A. D., Ahearn, T. S., Mustafa, N., Fox, H. C., & Whallety, L. J. (2012). Childhood socioeconomic status and adult brain size: Childhood socioeconmic status influences adult hippcampus size. Annals of Neurology, 71, 653–660. Stern, S. B., Smith, C. A., & Joon Jang, S. (1999). Urban families and adolescent mental health. Social Work Research, 23(1), 15–27. Stevens, C., Lauinger, B., & Neville, H. (2009). Differences in the neural mechanisms of selective attention in children from different socioeconomic backgrounds: An event-related brain potential study. Developmental Science, 12, 634–646. Strohschein, L. (2005). Household income histories and child mental health trajectories. Journal of Health and Social Behavior, 46(4), 359–375. Strully, K. W., Rehkopf, D. H., & Xuan, Z. (2010). Effects of prenatal poverty on infant health: State earned income tax credits and birth weight. American Sociological Review, 76, 534– 562. Takeuchi, D. T., Chung, R. C., Lin, K., Shen, H., Kurasaki, K., Chun, C., & Sue, S. (1998). Lifetime and twelve-month prevalence rates of major depressive episodes and dysthymia among Chinese Americans in los angeles. The American Journal of Psychiatry, 155(10),

1407–1414. Retrieved from http://search.proquest.com/docview/619368069? accountid=13158 Tee, J., & Kazantzis, N. (2011). Collaborative empiricism in cognitive therapy: A definition and theory for the relationship construct. Clinical Psychology: Science and Practice, 18(1), 47–61. Thornberry, T. P., Freeman-Gallant, A., Lizotte, A. J., Krohn, M. D., & Smith, C. A. (2003). Linked lives: The intergenerational transmission of antisocial behavior. Journal of Abnormal Child Psychology, 31(2), 171–184. doi: 10.1023/A:1022574208366 Tilghman-Osborne, C., Cole, D. A., & Felton, J. W. (2012). Inappropriate and excessive guilt: Instrument validation and developmental differences in relation to depression. Journal of Abnormal Child Psychology, 40(4), 607–620. doi: 10.1007/s10802–011–9591–6 Tolan, P. H., Gorman-Smith, D., Henry, D., Chung, K., & Hunt, M. (1998). The relation of patterns of coping of inner-city youth to social competence and psychopathology symptoms. Journal of Research on Adolescence, 12, 423–449. Tolan, P. H., Gorman-Smith, D., Henry, D., & Schoeny, M. (2009). The benefits of booster interventions: Evidence from a family-focused prevention program. Prevention Science, 10(4), 287–297. Tottenham, N., & Sheridan, M. (2010). A review of adversity, the amygdala and the hippocampus: A consideration of developmental timing. Frontiers of Neuroscience, 3, 1–18. Tracy, M., Zimmerman, F., Galea, S., McCauley, E., & Vander Stoep, A. (2008) Socioeconomic status and depressive symptoms in 11–13-year-olds: The role of stressful life events, family disruption, and neighborhood characteristics. Psychiatric Research, 42(14), 1163–1175. Twenge, J. M., & Nolen-Hoeksema, S. (2002). Age, gender, race, socioeconomic status, and birth cohort difference on the children's depression inventory: A meta-analysis. Journal of Abnormal Psychology, 111(4), 578–588. doi: 10.1037/0021–843X.111.4.578 U.S. Census Bureau, 2012. The Research Supplemental Poverty Measure: 2012. Current Population Reports. http://www.census.gov/prod/2013pubs/p60-247.pdf van den Bos, K., & Maas, M. (2009). On the psychology of the belief in a just world: Exploring experiential and rationalistic paths to victim blaming. Personality and Social Psychology Bulletin, 35(12), 1567–1578. Veenstra, R., Lindenberg, S., Oldehinkel, A. J., DeWinter, A. F., & Ormel, J. (2006). Temperament, environment, and antisocial behavior in a population sample of preadolescent boys and girls. International Journal of Behavioral Development, 30, 422–432. Wadsworth, M. E., & Achenbach, T. M. (2005). Explaining the link between low

socioeconomic status and psychopathology: Testing two mechanisms of the social causation hypothesis. Journal of Consulting and Clinical Psychology, 73, 1146–1153. Wadsworth, M. E., & Berger, L. E. (2006). Adolescent coping with poverty-related family stress: Prospective predictors of coping and psychological symptoms. Journal of Youth and Adolescence, 35, 57–70. Wadsworth, M. E., & Compas, B. E. (2002). Coping with family conflict and economic strain: The adolescent perspective. Journal of Research on Adolescence, 12, 243–274. Wadsworth, M. E., Raviv, T., Compas, B. E., & Connor-Smith, J. K. (2005). Parent and adolescent responses to poverty-related stress: Tests of mediated and moderated coping models. Journal of Child and Family Studies, 14, 283–298. Wadsworth, M. E., & Santiago, C. D. (2008). Risk and resiliency processes in ethnically diverse families in poverty. Journal of Family Psychology, 22, 399–410. Wadsworth, M. E., Raviv, T., Reinhard, C., Wolff, B., Santiago, C., & Einhorn, L. (2008). An indirect effects model of the association between poverty and child functioning: The role of children's poverty-related stress. Journal of Loss and Trauma, 13(2–3), 156–185. Wadsworth, M. E., Rindlaub, L., Hurwich-Reiss, E., Rienks, S., Bianco, H., & Markman, H. J. (2013). A longitudinal examination of the adaptation to poverty-related stress model: Predicting child and adolescent adjustment over time. Journal of Clinical Child & Adolescent Psychology, 42(5), 713–725. Wanless, S. B., McClelland, M. M., Tominey, S. L., & Acock, A. C. (2011). The influence of demographic risk factors on children's behavioral regulation in prekindergarten and kindergarten. Early Education and Development, 22, 461–488. Weikart, D. P., & Schweinhart, L. J. (1992). High/Scope preschool program outcomes. Preventing Antisocial Behaviour: Interventions from Birth Through Adolescence, The Guilford Press, New York. Werner, E. E., & Smith, R. S. (1982). Vulnerable but invincible. New York, NY: McGrawHill. Werner, E. E., & Smith, R. S. (2001). Journeys from childhood to midlife: Risk, resilience, and recovery. Cornell University Press. Whitbeck, L. B., Simons, R. L., Conger, R. D., Wickrama, K. A. S., Ackley, K. A., & Elder, G. H., Jr. (1997). The effects of parents' working conditions and family economic hardship on parenting behaviors and children's self efficacy. Social Psychology Quarterly, 60, 291–303. Wickrama, K. A.S., & Bryant, C. M. (2003). Community context of social resources and adolescent mental health. Journal of Marriage and Family, 65, 850–866. Wilkinson, R. G., & Pickett, K. E. (2009). The spirit level. London, UK: Allen Lane.

Williams, D. R., Mohammed, S. A., Leavell, J., & Collins, C. (2010). Race, socioeconomic status, and health: Complexities, ongoing challenges, and research opportunities. Annals of the New York Academy of Sciences, 1186, 69–101. Wills, T. A., McNamara, G., & Vaccaro, D. (1995). Parental education related to adolescentcoping and substance use: Development of a mediation model. Health Psychology, 14, 464– 478. Wilson, D., Foster, J., Anderson, S., & Mance, G. (2009). Racial socialization's moderating effect between poverty stress and psychological symptoms for African American youth. Journal of Black Psychology, 35, 102–124. Wiesner, M., & Windle, M. (2004). Assessing covariates of adolescent delinquency trajectories: A latent growth mixture modeling approach. Journal of Youth and Adolescence, 33(5), 431–442. Wodtke, G. T., Harding, D. J., & Elwert, F. (2011). Neighborhood effects in temporal perspective: The impact of long-term exposure to concentrated disadvantage on high school graduation. American Sociological Review, 76(5), 713–736. doi: 10.1177/0003122411420816 Wolf, S., Aber, J. L., & Morris, P. A. (2013). Drawing on psychological theory to understand and improve antipoverty policies: The case of conditional cash transfers. Psychology, Public Policy, and Law, 19(1), 3–14. doi: 10.1037/a0029498 Wolff, B. C., Santiago, C. D., & Wadsworth, M. E. (2009). Poverty and involuntary engagement stress responses: Examining the link to anxiety and aggression with low-income families. Anxiety, Stress, and Coping, 22, 309–325. Wolff, B. C., Wadsworth, M. E., & Santiago, C. D. (2011). Family poverty, stress, and coping. In R. J. R. Levesque (Ed.), Encylopedia of adolescence (pp. 941–951). New York, NY: Springer. Wray, L. A., Alwin, D. F., & McCammon, R. J. (2005). Social status and risky health behaviors: Results from the health and retirement study. Journals of Gerontology: Series B: Psychological Sciences and Social Sciences, 60B(2), 85–92. doi: 10.1093/geronb/60.Special_Issue_2.S85 Wyman, P., Cowen, E., Work, W., & Parker, G. (1991). Developmental and family milieu correlates of resilience in urban children who have experienced major life stress. American Journal of Community Psychology, 19, 405–426. Yanagisawa, K., Masui, K., Furutani, K., Nomura, M., Yoshida, H., & Ura, M. (2012). Family socioeconomic status modulates the coping-related neural response of offspring. Scan, 7, 1–6. Yoshikawa, H., Aber, J. L., & Beardslee, W. R. (2012). The effects of poverty on the mental, emotional, and behavioral health of children and youth: Implications for prevention. American

Psychologist, 67(4), 272–284. doi: 10.1037/a0028015 Youngstrom, E., Weist, M. D., & Albus, K. E. (2003). Exploring violence exposure, stress, protective factors and behavioral problems among inner-city youth. American Journal of Community Psychology, 32, 115–129. Zaslow, M. J., Moore, K. A., Brooks, J. L., Morris, P. A., & et al. (2002). Experimental studies of welfare reform and children. Future of Children, 12(1), 78–95. Retrieved from http://search.proquest.com/docview/220178316?accountid=13158

Chapter 5 Determinants of Parenting Marc H. Bornstein This chapter summarizes selected aspects of my research, and portions of the text have appeared in previous scientific publications cited in the references. This research was supported by the Intramural Research Program of the National Institutes of Health, Eunice Kennedy Shriver National Institute of Child Health and Human Development. I thank L. Henry for superb assistance. Address correspondence to: Marc H. Bornstein, Child and Family Research, Eunice Kennedy Shriver National Institute of Child Health and Human Development, National Institutes of Health, Suite 8030, 6705 Rockledge Drive, Bethesda MD 20892–7971, U.S.A. E-mail: [email protected]. INTRODUCTION PARENTING AND PARENTS Parenting Cognitions and Practices Parents Summary SOME METHODOLOGICAL CONSIDERATIONS AND FUTURE DIRECTIONS DETERMINANTS OF PARENTING DETERMINANTS IN THE PARENT Biological Characteristics Psychological Characteristics Summary DETERMINANTS IN THE CHILD Biological Characteristics Psychological Characteristics Summary DETERMINANTS IN THE CONTEXT Proximal Contexts Social Group Contexts Distal Contexts Summary

TRANSLATIONAL IMPLICATIONS CONCLUSIONS REFERENCES

Introduction Each day approximately three-quarters of a million adults around the world experience the pleasures and profits as well as the trials and tribulations of becoming a new parent. Each reader of this page has had the experience of being parented, and many relive the experience when they parent their own children. Evolutionary theory distinguishes between bringing a new individual into the world and caring for an existing individual, childbearing versus child caring (Bjorklund & Ellis, 2014; Bjorklund, Yunger, & Pellegrini, 2002). Whereas species lower in the phylogenetic hierarchy are principally childbearers, mammals such as Homo sapiens tend to be devoted child carers perhaps because the investment in a small number of progeny is great or, altricial, newborn human children are totally dependent on their parents. Human parents are fundamentally invested in their children: their survival, their socialization, and their education. Parents are tasked with preparing children for the physical, economic, and psychosocial situations that are characteristic of the environment and culture in which children must survive and, it is hoped, thrive (Benedict, 1938; Bornstein, 1991, 2012; LeVine, 2003). In consequence, understanding parenting is becoming a high-profile endeavor, and understanding the determinants of healthy and typical as well as unhealthy and atypical parenting is of increasing and compelling interest as well. Although parenting remains a somewhat mystifying subject about which almost everyone has opinions but about which few people agree, parenting study is enjoying popularity today as never before, and consequently a surprising amount of solid science is accumulating about parenting and its determinants. Adults already know (or think they know) something about parenting by the time they first become parents. Indeed, human beings appear to possess some intuitive knowledge about parenting, and certain characteristics of parenting may be wired into our biological makeup. For example, people almost everywhere speak to babies even though they know that babies cannot understand language, and people even speak to babies spontaneously in a speech register specially reserved for them. However, human beings also acquire important knowledge of what it means to parent children through generational, social, and media images of parenting, children, and family life, knowledge that plays a significant role in helping people formulate their parenting cognitions and guide their parenting practices. For example, parents from different cultures differ in their opinions about all manner of caregiving-related issues, such as the relevance of diverse competencies specific for their children's successful adjustment, the ages they expect children to reach different developmental milestones, and so forth. Direct experiences with children and self-constructed parenting also help formulate parenting attitudes and actions. What forces affect when and how parents parent? This question of determinants goes to the heart of much of contemporary parenting science. It is the subject matter of this chapter. In

consequence, the chapter does not dwell on parenting as a phase of adult development or parenting as an instrumental activity vis-à-vis children and their development. Of course, parents are concerned about themselves as parents, and parents are concerned about the everyday well-being of their children as well as their children's long-term health and growth. Nor does the chapter address the origins of parenting studies, theories of parenting, or arguments for and against parenting effects. For an introduction and overview of these important parenting issues, see the companion chapter in Bornstein (2015). Rather, this chapter surveys central issues concerning determinants of parenting, both normal and typical as well as abnormal and atypical (Cicchetti & Toth, 2009); it pursues the following course. The first section of the chapter briefly overviews, in seriatim, cognitions and practices that principally instantiate parenting and then parents, the principal actors in the human drama of caregiving. As a transition to the main focus on the multiple determinants of parenting, the chapter next alerts the reader to a series of methodological concerns useful in understanding and unpackaging determinants of parenting and points to some gaps in the literature that suggest future directions of research. The chapter then turns to the main substance—a review of the determinants of parenting in parents, children, and contexts. The chapter then addresses translational implications and concludes.

Parenting and Parents As judged by psychoanalysis, ethology, psychology, and neuroscience, parents engage in a peculiar kind of life's work: Parenting is a delicate blend of altruism, prosociality, devotion, and selflessness, marked by constantly challenging demands, changing and ambiguous criteria, and all too frequent evaluations. Parenting entails both affective components—in terms of commitment, empathy, and positive regard for children—and cognitive components—the how, what, when, and why of caring for children. So, thinking about the determinants of parenting demands multilevel considerations of psychobiology, cognition, emotion, and sociality as well as multidisciplinary considerations of biology, psychology, sociology, and anthropology. According to a nationwide survey conducted by the National Center for Children, Toddlers, and Families, more than 90% of parents say that, when they had their first child, they not only felt “in love” with their baby but were personally happier than ever before in their lives (Civitas Initiative, Zero to Three, & Brio Corporation, 2000). Parents report relatively higher levels of happiness, positive emotion, and meaning in life than do nonparents (Nelson, Kushlev, English, Dunn, & Lyubomirsky, 2013). Parents find interest and derive considerable and continuing personal reward in their relationships and activities with their children. They report that spending time with their children, especially in recreation or educational child care, ranks among their most enjoyable activities (Krueger, Kahneman, Schkade, Schwarz, & Stone, 2009; Nelson et al., 2013). And they are doing more of it: Bianchi (2000), Sayer, Bianchi, and Robinson (2004), and Aguiar and Hurst (2007) concur that adults in the United States (as elsewhere in the world; Gauthier, Smeeding, & Furstenburg, 2004) are spending more time with their children today than in the past. Notably, mesocortico-limbic dopamine is involved in both reward-related learning and mammalian caregiving (Insel, 2003).

For many, parenthood enhances psychological development, self-confidence, and a sense of well-being. Parenting translates into a constellation of new trusts, provides opportunities for enrichment and learning, and unveils a panorama on the “larger picture” of life. Parenthood also gives adults ample occasion to confront new challenges and to test and display their competencies (Crittenden, 2004). Markus, Cross, and Wurf (1990) learned that feelings of competence as a parent constitute a highly common aspect of the self as desired by adults. More commonly, however, parenting is identified with its instrumental character, as acts of providing for and supporting the biological, physical, intellectual, emotional, and social development of progeny. Human children cannot and do not grow up as solitary individuals, so parents want to know how best to take advantage of the opportunities parenting affords and cope with parenting's unrelenting demands. Some consider this instrumental construal of parenting a lifelong 24/7/365 job. Parenthood is a major phase and constituent ingredient of mature and generative adulthood, and parenthood is central to childhood, to child development, and to society's long-term investment in children. Thus, we are motivated to know about the meaning and importance of parenthood and parenting and about how parenting develops. First, however, we address two introductory questions. How is parenting expressed, and who parents?

Parenting Cognitions and Practices Parenting is instantiated in cognitions and practices, what parents think and what they do. Child rearing entails all sorts of mentation and consists of multiple kinds of interactions of parent, child, and context; parents possess attitudes, expectations, and attributions, and they naturally engage their children in a range of diverse activities and do not only or necessarily behave in uniform ways. Parenting appears to be a social construction that judiciously combines personological consistency with situational specificity. Instead of broad styles, it may be more appropriate and valid to conceive of parenting as fine-grained domains of practices and cognitions. Parenting is therefore multidimensional, modular, and specific (Boivin et al., 2005; Bornstein, 2002a, 2006a; Bornstin et al., 2008; Bradley & Caldwell, 1995; E. Skinner, Johnson, & Snyder, 2005) and so supports the identification and empirical focus on relatively independent cognitions and practices. When the chapter turns to determinants of parenting, it is concerned with the determinants of variegated cognitions and practices. Parenting cognitions have long held a popular place in the study of parenting, and they span a wide array of goals, attitudes, expectations, perceptions, attributions, and actual knowledge about parenthood, about one's own parenting, about childhood, and about one's own child(ren). Cognitions also form a framework within which parents perceive and interpret their children's behaviors. Parenting practices give expression to parenting cognitions and constitute a large measure of children's worldly experience. The contents of human parent-child interactions are dynamic and varied and include multiple obligatory as well as discretionary activities. Parents nurture and protect children, but they also guide children in understanding and expressing proper feelings and emotions, educate children as to actions that are acceptable for the stage of childhood they occupy, and prepare children for adaptation to a wider range of life roles and contexts they will encounter as they grow. Take protectiveness as an expression of parental

warmth (i.e., affection, nurturance, and responsiveness). The developmental literature historically treated protectiveness as a unidimensional construct (Parker, 1983; Sargent, 1983). Besides trying to protect their children from physical harm, however, parents also engage in diverse behaviors to protect their children from psychological and social harm (Power, 2004). Not unexpectedly, different kinds of protectiveness show different relations with different domains of child adjustment (Power & Hill, 2008). So parent protectiveness has proved to be better conceptualized as illustrative of the multidimensional, modular, and specific nature of parenting, and the effectiveness of parent protectiveness is better analyzed with respect to its different facets and effects. In practice, of course, parent-child interaction is dynamic, intricate, and meshed, and parents regularly engage in combinations of cognitions and practices. Taken as a totality, parenting cognitions and practices constitute a varied and demanding task set, and adults differ considerably in terms of how they esteem various components of the caregiving repertoire as well as in how successful they are in executing different components. The attitudes and actions of children's parents, and the environments of caregiving they create, matter in that they shape children and the course of children's academic, emotional, behavioral, and social development (Bates, Schermerhorn, & Petersen, 2012; Belsky & de Haan, 2011; Bornstein, 2015; Collins, Maccoby, Steinberg, Hetherington, & Bornstein, 2000; Domitrovich & Bierman, 2001; Goodnow, 2002; Holden & Buck, 2002; Maccoby, 1992; Roksa & Potter, 2011; Sigel & McGillicuddy-De Lisi, 2002; Stormshak, Bierman, McMahon, & Lengua, 2000; Vandell, 2000). We know this from a long history and wide variety of convergent approaches from correlational designs to experiments to interventions. For example, Patterson and Forgatch (1995) reported substantial correlations between parents' disciplinary and monitoring practices and children's negative and coercive behavior both at home and in out-of-home contexts, and Chilcoat and Anthony (1996) reported significant increases in risk of adolescents' drug sampling for every unit of decrease in parental monitoring after partialing age, gender, and ethnic status. The foregoing studies, although based on correlations, illustrate associations between parenting and child development that transcend a common context and obtain above common-cause third-variable controls (Ganiban, Saudino, Ulbricht, Neiderhise, & Reiss, 2008; Knafo & Jaffee, 2013; Price & Jaffee, 2008). Longitudinal associations, adoptive parent–adopted child associations, and other emerging conservative strategies that supplant simple zero-order correlations constitute additional design improvements on cross-sectional and genetically related parent-child approaches. For example, higher levels of warm and sensitive parenting predict later decreases in child negative reactivity, even controlling for initial levels (Bates et al., 2012), and higher levels of harsh and controlling parenting predict increases in child negative reactivity, even controlling for initial temperament (Belsky, Fish, & Isabella, 1991; Braungart-Rieker, Hill-Soderlund, & Karrass, 2010). Using an adoption design that differentiates heritable (from biological parent) and environmental (from adoptive parents) effects on the adoptee's behavior, Ge et al. (1996) found that adoptive parents' behavior was associated with the psychiatric status of the adoptee's biological parents and that this association was mediated by the adoptee's behavior: Adoptees born to parents with psychopathologies were more hostile and antisocial, and their adoptive parents used more harsh discipline and were less nurturing and involved (see also

O'Connor, Deater-Deckard, Fulker, Rutter, & Plomin, 1998). Apparently having genes of parents with psychopathologies affects adoptees' behavior and thus adoptive parents' parenting as well. Adoptive parent educational achievement is associated with adopted children's schooling (Bjoerklund, Lindahl, & Plug, 2006). Rice et al. (2010) employed a prenatal crossfostering design, including pregnant mothers who were related or unrelated to their child as a result of in vitro fertilization and surrogacy, to disentangle inherited and environmental influences. Their logic was: If links between prenatal stress and offspring outcome are environmental, the association should be observed in unrelated as well as related mother-child pairs. Associations between prenatal stress and offspring birthweight, gestational age, and antisocial behavior were seen in both related and unrelated mother-offspring pairs, consistent with environmental links being determinative. The association between prenatal stress and offspring anxiety in related and unrelated groups appeared to be due to current maternal anxiety/depression rather than prenatal stress. By contrast, the link between prenatal stress and offspring attention-deficit hyperactivity disorder (ADHD) was present only in related motheroffspring pairs and therefore was likely attributable to genetic factors. Experimental manipulations advance beyond parent-child correlations in attempts to document causal relations between parenting cognitions or practices and children's development. Animal experiments and natural experiments, designed experiments, and experimental interventions with human beings give evidence of parenting effects. Experiments with animal populations since Harlow's (1958) studies of monkeys raised with contrasting surrogate mothers provide invaluable and unique data about parenting because many informative manipulations are unethical with human beings. Variation in maternal care by rat dams (specifically arched-back nursing and licking and grooming) during the first postnatal week affects endocrine, behavioral, and even cognitive development of offspring: Pups of maternal dams show a decreased stress response, increased glucocorticoid receptor expression and synaptogenesis in the hippocampus, and enhanced spatial memory compared to pups of nonmaternal dams (Fish et al., 2004). The same effects also appear in cross-fostered pups (I. C. G. Weaver et al., 2004), demonstrating that they are not due to genetic inheritance. Notably, these endocrine, behavioral, and cognitive effects persist to maturity. Experiments that have randomly assigned human families to treatment versus control groups and that intervene with the parents but do not simultaneously treat the children are rare (for obvious reasons), but several have shown that treatment can change parental thinking and action toward children in specified ways and, in consequence, alter child development. For example, intranasal administration of neuropeptides can reach the central nervous system, and oxytocin (OT) is known to play a key role in regulating social behavior and supporting the parent-infant bond (MacDonald & MacDonald, 2010). Naber, van IJzendoorn, Deschamps, van Engeland, and Bakermans-Kranenburg (2010) videorecorded fathers during a play session with their toddlers once following intranasal OT administration and then after intranasal placebo administration. Fathers increased their support for learning and exploration with respect for the child's autonomy following OT administration. Utilizing a double-blind, placebo-controlled crossover design, Weisman, Zagoory-Sharon, and Feldman (2012) observed fathers and their infants twice in a face-to-face still-face paradigm following

administration of OT or placebo. OT administration increased fathers' salivary OT and key parenting behaviors that support parent-infant bonding. Moreover, parallel increases were found in infant salivary OT and engagement behavior, including social gaze, exploration, and reciprocity. In other words, OT administration had parallel effects on the treated parent and untreated child. Interventions with parents have two construals. Near the end of this chapter, interventions are discussed for their translational implications to improve parenting clinically and to inform policy. Intervention trials also can be interpreted as experimental manipulations that test theoretical models of parenting effects (P. A. Cowan & Cowan, 2002; van Doesum, RiksenWalraven, Hosman, & Hoefnagels, 2008). An intervention designed to facilitate maternal responsiveness enhanced children's language skills (Landry, Smith, Swank, & Guttentag, 2008); a second to improve parents' behavior support found positive effects on key indicators of children's school readiness (Lunkenheimer et al., 2008); and a third to boost parents' positive behavior practices reduced problem behaviors in young children (Dishion et al., 2008). Parenting in one generation is consequential because it has manifest instrumental validity for the succeeding generation. Importantly, individual differences in parenting cognitions and practices tend to be stable, meaning that parenting assessed at one point can to a certain degree be assumed to reflect past as well as future parenting. Short-term “test-retest” reliabilities of cognitions (r = .74) and practices (r = .59) in parents tend to be high (Holden & Miller, 1999). Holden and Miller (1999) meta-analyzed the consistency of beliefs and behaviors of parents (mostly mothers) with both relative standing and mean level change in mind and considering three circumstances: comparisons across time, children, and situations. Consistency was found to depend on circumstance as well as parenting construct, child age, and methodological approach among other moderators. They arrived at diminishing summary levels of childrearing stability across children (r = .50), across time (r = .45), and across situations (r = .26) and increasing differences across children (d = .23), across time (d = .40), and across situations (d = .52) (see also Bornstein, Gini et al., 2006; Dallaire & Weinraub, 2005; Haltigan, Roisman, & Fraley, 2013; Maas, Vreeswijk, & van Bakel, 2013; Porter & Hsu, 2003, for supporting instance cases since the Holden and Miller meta-analysis).

Parents Now that the principal attitudes and actions in parenting have been identified, it is appropriate to turn to consider parenting's principal actors. Who parents children? The majority of children throughout the world grow up in family systems with more than one or two significant parenting figure guiding a child's upbringing and socialization, even if the research literature usually focuses on mothers. Mothers and fathers are not children's only caregiving agents. The conception of parenting as a set of cognitions and practices expands the discussion of determinants of parenting beyond biological parents to other related (siblings, grandparents) and unrelated (adoptive, foster, and alloparents) caregivers who may also “socially” parent children (Leon, 2002). This chapter is largely circumscribed to determinants of parenting in parents, although for completeness sake determinants of some nonparental caregiving are

briefly described. Mammals tend generally to be devoted child “carers” (Bjorklund et al., 2002), and almost all mammalian species are matrilocal, the norm in mammalian care falling to mothers (CluttonBrock, 1991; Wilson, 1975). On this account, maternal care is much more common than paternal care (in mammals, males provide care in less than 5% of species; Moller, 2003). Most observers agree that mothers (and female relatives) normally play the more central role in human child rearing (Barnard & Solchany, 2002; Civitas Initiative et al., 2000; Holden & Miller, 1999; Leiderman, Tulkin, & Rosenfeld, 1977; Weisner & Gallimore, 1977) and in many nonhuman mammalian species (Briga, Pen, & Wright, 2012). The maternal role is better articulated and defined than is the paternal role, and in their own upbringing, women are generally afforded more opportunities to acquire and practice skills that are child centered. Mothers are primarily responsible for home and family, and they are believed to bear a heavier psychic burden in parenting (Barnard & Solchany, 2002; Calzada, Eyberg, Rich, & Querido, 2004; Metsäpelto & Pulkkinen, 2003; Verhoeven, Junger, Van Aken, Deković, & Van Aken, 2007). Even among species where males show considerable parental altruism, they commonly do less caregiving work than females (Wilson, 1975). One finds many more nonresident dads than moms (Sorensen, 1997), for example. Homo sapiens is an alloparental species, and cooperatively breeding vertebrates (like human beings) often enlist “helpers” who take care of immature offspring who are not their own within the social group (Hrdy, 2009). Among humans, siblings sometimes care for children (Zukow-Goldring, 2002); grandparents often assume central roles in child caregiving (P. K. Smith & Drew, 2002); and numbers of young children around the world normatively participate in diverse types of nonparental nonfamilial care (Clarke-Stewart & Allhusen, 2002). Grandparental help may be indirect or direct. Increased life expectancy, decreased family size, more maternal employment, and the rise of single-parent families have conspired to increase the potential for grandparents (and others) to play greater parts in the lives of their grandchildren (Arber & Timonen, 2012; Dunifon, 2012; Kornhaber, 2002; Tanskanen & Rotkirch, 2014; Witkin, 2012). Approximately 43% of grandparents provide some child care on a regular basis (Lou & Chi, 2012; Stelle, Fruhauf, Orel, & Landry-Meyer, 2010), and in custodial grandparent families, children are reared solely by grandparents (as the result, e.g., of maternal incarceration or other parental problems; Poehlmann et al., 2008). Many tasks associated with child care can also be purchased in the marketplace. Today children enter that kind of care early, stay for longer periods of time, and change types of care often (Burchinal, Magnuson, Powell, & Hong, 2015). Beginning in the child's first year, about 50% of the children in the United States, for example, experience regular nonparental child care, and by the preschool years, more than 75% of children have experienced some type of child care. As a consequence of contemporary social and cultural forces (mentioned earlier), the use of child care services has burgeoned, and nonparental caregivers have assumed increasing responsibility for meeting children's developmental needs and preparing children for a future in society (Gottfried, Gottfried, & Bathurst, 2002). It has been argued that mothers and fathers and others do not necessarily share the same

“investments” in parenting (Geary, 2000), and human cultures distribute the tasks of child care in different ways (Leinaweaver, 2014). Children's parents and others divide the labor of caregiving, and the content, meaning, and effects of mother-child, father-child, and other-child relationships all have significances. Indeed, researchers have reckoned that, for example, mothers and fathers make independent contributions to children's development (A. Martin, Ryan, & Brooks-Gunn, 2007; Ryan, Martin, & Brooks-Gunn, 2006), children develop attachments to different people (Howes & Oldman, 2001; Howes & Spieker, 2008), and social relationship theory posits that multiple associations are important to children because they meet different developmental needs (Vandell, 2000). All said, it comes as little surprise that a more extensive body of information has developed about determinants of mothering than about fathering or other caregiving. However, the division of child care implies that developmental science eventually needs to consider determinants in mothers, fathers, as well as multiple other “parents.”

Summary Parenting expresses itself in both the beliefs parents hold and the behaviors they exhibit. Many conceptually separate parenting cognitions and practices have been discerned, and each is developmentally significant. Moreover, many people parent children. With parenting so demonstratively meaningful, we should ask not whether parents matter, but what makes a good parent. Parenting is important; therefore, the determinants of parenting are important. Parenting is not an activity people normally think of as being especially scientific. Like most things, however, better parenting benefits from greater knowledge and understanding, which in turn depend on theory and research. Happily, there is a science of parenting with developing theory and growing systematic research behind it. The contemporary parenting literature, however incomplete, now contains thousands of studies. This chapter of determinants of parenting is based on that consolidating science of parenting.

Some Methodological Considerations and Future Directions Measurement is an importunate challenge in contemporary parenting science (Bornstein, 2010a; Bornstein & Toole, 2010). Hurley, Huscroft-D'Angelo, Trout, Griffith, and Epstein (2014) reviewed the psychometric quality of parenting measures published from 1985 to 2009. Their initial search yielded 164 English-language measures, which was reduced to 25 that supplied some psychometric information. Considering 10 psychometric properties, out of the 25, 7 had no acceptable psychometric properties, 7 had only 1 or 2, 6 had 3 to 4, none had 5 to 6, and only 5 had strong psychometric properties in 7 or more. Only 5 measures provided norming information. The parenting literature is also encumbered by mixed findings, and empirically there is widespread lack of agreement regarding which assessment methodology may be most effective and valid to use, as observation, test, and interview and questionnaire methods have different

strengths and limitations (Bögels & van Melick, 2004; Collett, Gimpel, Greenson, & Gunderson, 2001; Dadds, Maujean, & Fraser, 2003; Gaylord, Kitzmann, & Coleman, 2003; O'Connor, 2002; Reitman, Rhode, Hupp, & Altobello, 2002; Rhoades & O'Leary, 2007). Moreover, which features of parenting are most vital to study presents a central, perennial, and sometimes vexing choice (Bornstein, 2013b). Related empirical concerns focus on the need for measurement invariance of assessments across parent gender, socioeconomic status (SES), and culture (Adamsons & Buehler, 2007; Furman & Lanthier, 2002; L. Huang et al., 2011; Locke & Prinz, 2002; Senese, Bornstein, Haynes, Rossi, & Venuti, 2012; Stolz, Barber, & Olsen, 2005). Additional challenges for the future study of determinants are to acknowledge and rectify the fact that that, up to now, the rapidly developing discipline of parenting research has focused too narrowly on determinants of mothering to the near exclusion of other caregivers and of multiple family relationships; on selected determinants of parenting, such as SES, to the near exclusion of others, such as religion; on determinants in normative nuclear families, when the world is populated with a bewildering panoply of family compositions (Ganong, Coleman, & Russell, 2015; Perrin, Siegel, & the Committee on Psychological Aspects of Child and Family Health, 2013; Shah, Zao, & Ali, 2011); and on determinants of parenting in the minority (Western, educated, industrialized, rich, and democratic) developed world rather than the majority developing world (Bornstein, 2010b; Henrich, Heine, & Norenzayan, 2010; Tomlinson, Bornstein, Marlow, & Swartz, 2014). Future cutting-edge research designs will also account for assortative mating, control one parent in analyzing determinants on the other parent, and other areas. Another pervasive critique of contemporary parenting science is that research in the area has not yet adequately addressed determinants of the nuanced complexities of everyday parenting, where the rubber meets the road. Parents multitask; parenting is more than feeding and teaching and being emotionally available. Parenting is planning and organizing (arranging playdates, identifying after-school activities, finding pediatricians) that consumes mental and physical energy and time. Parenting is disambiguating moment-to-moment novel, complex, dynamic, uncertain information associated with children and child rearing. Parenting is expecting of oneself and being expected by others to caregive consistently, appropriately, and effectively. Parenting is being highly motivated to succeed at these many assignments. Using nonparticipant unobtrusive naturalistic observational methods, Radesky and colleagues (2014) captured frequency, duration, and modality of mobile device use versus caregiving of people tending young children eating in fast food restaurants. Caregivers who were highly absorbed with their devices also responded harshly to child misbehavior. Consumers of the determinants of parenting literature should consider that parenting also affects each putative determinant, and untangling direction of effects is a persistent vexing concern. This field needs to move beyond single-point-in-time cross-sectional determinant-ofparenting studies that leave the direction of parenting determinants ambiguous. Strong evidence that a given factor “determines” parenting would be a Time 1 (T1) determinant–Time 2 (T2) parenting link that (good) controls for the T1 level of the parenting variable in question and (better) also controls for the concurrent T2 association between the putative determinant and the parenting outcome. Such designs are supported by, for example, a partial correlation

between T1 determinant and T2 parenting, controlling for T1 parenting or a cross-lag correlation in which the T1 determinant predicts the part of the variance in T2 parenting that remains after controlling for T1 parenting and its concurrent link to the T2 determinant or a comparable regression formulation. (Simple longitudinal precedence is only suggestive of causality; Reiss, 1995.) When is a determinant effective? Consider three possible temporal models of the effectiveness of some determinant of parenting. As described, an antecedent determinant of parenting might be effective and endure, independent of later experience with the determinant and of any earlier contribution of the parenting variable. This antecedent determinant model is consonant with a sensitive period interpretation of determinant effects on parenting (Bornstein, 1989). Alternatively, a determinant could exert a unique influence on parenting only at current points in time that override the effects of antecedent influences. Or, in a cumulative/additive/stable model, a determinant at any one time might not necessarily affect parenting, but meaningful longitudinal relations need to be structured by the determinant continually repeating and aggregating through time. To date, theoretical explanations for specific pathways and directions of associations are weak, and methodologically rigorous studies (e.g., using longitudinal data in conjunction with independent assessments of determinants) continue to be the exception. The issue of determinants raises the important questions of multicausality. On one hand, many “causes” of parenting naturally covary, so multicollinearity is an issue that merits close attention. For example, youth is associated with less than optimal parenting, perhaps because young parents themselves are still developing executive functions or a sense of self, but people who have children early in their life career are also more likely to have low education and low SES, which adversely affect their parenting. Bandura (1997) defined self-efficacy as personal judgments about one's ability to execute a future course of action and to encompass feelings of competence about one's ability to perform a role or task. Self-efficacy has come to occupy a central role in parenting science and is linked with positive parenting. In Bandura's analysis, efficacy is influenced by a number of factors, including task difficulty, effort expenditure, vicarious experiences, physiological and mood states, social or verbal persuasion, and outcome expectancies that have been formed by past successes and failures. Self-efficacy as much as any construct exemplifies multideterminacy in parenting. As we shall see, Bandura underestimates the case, as a slew of other factors, from depression to child care experience to social support, likewise co-construct self-efficacy in parents. To identify any one determinant of parenting, it is necessary to covary third-variable confounders. On the other hand, it is certain that parenting has many determinants, and it is illogical and nonscientific to assert the preeminence of one cause over another when each in its own way contributes to some effect. Within complex developmental systems, it is unlikely that any single factor can be expected to account for substantial amounts of variation. Parenting effects are also conditional and not absolute (i.e., true for all parents under all conditions). More complex conceptualizations that incorporate larger numbers of influential variables will likely explain parenting better than simpler ones with fewer variables. The constructive enterprise is really to understand how all relevant forces might work in concert. Parenting is a multilevel phenomenon and will be better understood eventually by employing evolutionary, biological,

comparative, behavioral, and cultural perspectives. By 1997, Holden reported that more than 30 variables had been identified to empirically influence parenting. The literature concerning endogenous and exogenous sources of influence on parenting has since burgeoned further, but antecedents have typically been studied in isolation, and few investigations evaluate multiple influences simultaneously. Thus, the overlap of different antecedents vis-à-vis the unique contribution that any one may make to parenting remains essentially unexplored. With this limitation in mind, family systems theorists have emphasized the importance of considering the possible independence and interdependence of multiple organismic, environmental, and experiential determinants of parenting. Various models of action might obtain: It could be that some determinants compete with each other, others may be additive, and still others interact (Holden & Miller, 1999). Furthermore, effect sizes for any one determinant-parenting relation, or for contingencies that operate across different components of the family system, might be small in magnitude. This is not a statement about the unimportance of small effects (see Bornstein, 2014) but rather of the need to build models of parenting that focus on the ways small effects might combine. The number and complexity of determinants have implications for the expression of parenting. Holden and Miller (1999) pointed out that, if parenting is determined by a small number of consistent or operative variables, parenting might be characterized by great similarity. If, however, parenting is determined by a large number of influences or variables that change, parenting might be less consistent. The still-inadequate situation of contemporary parenting research is therefore ripe for revolution and advancement. In the meantime, the account of determinants of parenting that follows must be understood in light of extant literature, and in weighing the ensuing survey, the reader is best informed by a clear understanding of these several constraints and shortfalls as well as others that arise as successive topics are discussed.

Determinants of Parenting For heuristic purposes, the considerations of parent, child, and contextual determinants of parenting in this chapter treat each factor as a determinant, and each has been investigated as such. However, in an organized, linked, multicausal system, any element has the potential to moderate both the structure and the impact of other elements in the system. Suppose (following Bronfenbrenner & Morris, 2006) that parenting is the joint product of four defining properties: Process × Person × Context × Time. Process refers to dynamic interactions that the parent experiences. In considering child development, Bronfenbrenner and Morris (2006) gave central attention to “proximal processes,” which they posited to be most important. Applied here, such processes would consist of the parent's experiences within microsystems and include social interactions with others and engagement with particular materials in a setting. Characteristics and qualities of the parent qua person, factors such as age, gender, personality, and competencies, are posited to determine parenting.

Context refers to nested micro-, meso-, exo-, macro-, and chronosystems which Bronfenbrenner (1979) identified and distinguished in conceptualizing different spheres of influence on an individual. To paraphrase, he defined the microsystem as the pattern of activities, roles, and interpersonal relationships experienced by the parent in face-to-face settings with particular physical and material features. The microsystem also can encompass mass media (television, movies, or the Internet), given that these sources affect a parent's functioning. For example, based on a national U.S. sample of 1,000 parents living with at least one target child 6 to 17 years old whom Romer and his colleagues (2014) exposed to sequences of three pairs of short scenes with either violent or sexual content from popular movies, parents became progressively desensitized to both violence and sex in movies, which may contribute to their increasing acceptance of both types of content for young children. The parent microsystem is nested under the mesosystem, defined in paraphrase as linkages and processes that take place between two or more settings containing the parent. A mesosystem is in essence a system of microsystems. For example, mothers are likely influenced by information they receive from their support networks (McConnell, Breitkreuz, & Savage, 2011) or when attending a school event affects their parenting (McIntosh, Lyon, Carlson, Everette, & Loera, 2008). The exosystem encompasses linkages between features of the environment the parent does not directly encounter but that guide parenting through their effects on the microsystem and mesosystem. For example, parents' use of computer technology to telecommute rather than work at their workplace illustrates an exosystem setting effect on parenting (Golden, 2012). At the outermost macrosystem level are overarching patterns of beliefs, values, customs, and living conditions. Bronfenbrenner (1979) highlighted the importance of macrosystem constituents, such as religion, social class, and culture. Cutting across these four levels is the dimension of time, the chronosystem, which includes minute-byminute exposure to various processes, the periodicity of episodes over broader intervals of days or weeks, and even longer life course and historical eras. The chronosystem can be viewed in terms of the comparative salience of proximal processes occurring at different ages (e.g., critical or sensitive periods; Bornstein, 1989) or in terms of persons encountering different macrosystem events at different ages (Elder & Shanahan, 2006). Bioecological theory provides a rich and generative framework for drawing attention to important active ingredients and features of different determinants and in formulating ways to conceptualize relations between different determinants. A crucial point is that the meso-, exo-, macro-, and chronosystems are not separate from the parent's immediate microsystem. Rather, each system permeates those inferior to it. Moreover, it is important to recognize the bidirectional nature of system linkages. For example, the interplay between parent work conditions and family functioning, a usual exosystem analysis, involves assessments of work→family (work-to-family spillover), where work demands infringe or enhance parent functioning (Hsueh & Yoshikawa, 2007; Krapf, Ursprung, & Zimmerman, 2014), as well as family→work (family-to-work spillover), where parent responsibilities infringe or enhance work performance (Buehler & O'Brien, 2011; Krapf et al., 2014). This system structure is best viewed as hierarchical in nature, with lower-order elements nested within higher-order elements and no one level predominant (Ungar, Ghazinour, & Richter, 2013). Here I adopt this

scheme to model the determinants of parenting. A further important consequence of multiple linkages across system levels reinforces the probabilistic nature of parenting, given that the impact of influences from one level can be moderated by characteristics at other linked levels. A vital step on the path to fully understanding parenting is to evaluate the many forces that structure it. The origins of individual variation in maternal, paternal, and alloparental caregiving, whether of cognitions or practices, are extremely complex, and it has long been explicitly acknowledged that parenting is multiply determined. Contexts (such as evolution, history, ecology; culture, ethnicity, religion, SES; neighborhood, networks, family, situation), children's structural and psychological characteristics, and parents' own biological and psychological characteristics work in tandem to construct a parent. Consistent with a relational developmental bioecological orientation (Belsky, 1984; Bornstein, 2006a; Bornstein et al., 2007; Bronfenbrenner & Morris, 2006; Farnfield, 2008), this vast array of determinants can be grouped roughly (following the logic of hierarchical regression) according to their proximity to the parent from proximal to distal: (a) intrapsychic and intrapersonal characteristics of parents, (b) actual or perceived characteristics of children, and (c) contextual characteristics. A thorough and thoroughly up-to-date meta-analysis sought to quantitatively synthesize what twin and adoption studies have to say about the etiology of parenting (Klahr & Burt, 2014). Unsurprisingly, this meticulous attempt to definitively catalog genetic and environmental sources of parenting identified effects attributable to parent, child, and context. Reasons of space constrain a full accounting of all possibilities, and so this exposition of the determinants of parenting in the three main sources is illustrative rather than exhaustive.

Determinants in the Parent Because securing the survival of offspring is fundamental to evolutionary pressure, it is likely that specific biological and psychological characteristics evolved in the service of parenting. Expressions of parenting have been shown to reflect biological and psychological characteristics that equip parents (an even nonparents) to interpret and respond to the kaleidoscopic requirements and responsibilities of parenting. Even if most parents face the formidable challenges of parenthood with some degree of naiveté, parents do not meet the task totally unprepared.

Biological Characteristics Genetic endowment, neurohormonal activity, and central nervous system structure and function constitute some of the central biological characteristics identified to condition parenting. Genetics Behavior genetics (BG) seeks to understand biological sources of variation in human characteristics (Plomin, DeFries, McClearn, & McGuffin, 2008). By studying individuals of varying genetic relatedness (identical and fraternal twins, biological and adopted siblings who share or do not share the same experiences), behavioral geneticists attempt to estimate the

amount of variation (heritability; h2) in characteristics that can be explained by genetic endowment. Heritability denotes inheritance of DNA and refers to the main effect of genetics on individual differences but does not necessarily imply genetic determinism (for discussions of interpretations and misinterpretations of heritability, see Plomin et al., 2008; and Visscher, Hill, & Wray, 2008). BG assumes that sources of variation in a parent characteristic can be separated into independent genetic (G) and environmental (E) components that together (with error variance) add to l00% of the variance in a characteristic. G effects are additive (A) and dominant (D), and E effects are shared (C) and nonshared plus error (E). (E is not usually measured directly but estimated as the residual variance not accounted for by G; see Caspi, Taylor, Moffitt, & Plomin, 2000.) From a BG perspective, parenting is just another phenotype that is a reflection of nature and nurture and not an eventuality of “pure” nurture (McGuire, 2003). Genetic contributions to parenting are conceptualized as evidence of genotype-environment correlations (rGE). Three types of rGE processes have been articulated: Passive rGE processes occur when family members share both environments and genes; evocative rGE processes refer to the extent to which genetically linked traits in individuals elicit responses from others; and active rGE processes refer to the extent to which people seek out environments that are correlated with their genetically linked traits. Studies of the heritability of parenting usually focus on passive and evocative processes because active rGE processes are difficult to detect in the absence of longitudinal data (McGuire, Segal, & Hershberger, 2012). Lytton and Gallagher (2002) and McGuire (2003) introduced and discussed BG research on the heritability of parenting beliefs and behaviors. Early BG studies focused on demonstrating that some parenting cognition or practice is heritable; for example, parents' warmth (Neiderhiser et al., 2004). Current studies investigate moderators and mediators of genetic and environmental contributions to parenting (McGuire et al., 2012). Kendler and Baker (2007) meta-analyzed child- and parent-based BG studies of contributions to parenting and found that weighted heritabilities from 12 studies that were child-based were highest for parental warmth (34–37%), then protectiveness (20–26%), and last control (12–17%), and 10 parent-based studies showed higher heritabilities for parental warmth (35%) compared to other parenting dimensions (19–23%). A succeeding review (Klahr & Burt, 2014) arrived at similar conclusions with respect to heritabilities for different dimensions of parenting in mothers and fathers based on child and parent adoption and twin designs. A full consideration of the BG literature and the controversies it raises is beyond the scope of this chapter. Suffice it to say that empirical and philosophical points have been argued to vitiate many central BG claims (Charney, 2012; Collins et al., 2000; Elman et al., 1996; Gottlieb, 1995; Rose, 1995; Turkheimer, 1998). For example, heritability estimates are often indeterminate and variable (Klahr & Burt, 2014), as estimates fluctuate depending on the dependent construct, source of information, and other methodological factors; genes may function differently in different environments and at different times (Naumova, Lee, Rychov, Vlasova, & Grigorenko, 2013; Szyf & Bick, 2013); and meta-analyses reveal that heredity rarely accounts for as much as 50% of the variation among individuals in a particular population (Klahr & Burt, 2014; McCartney, Harris, & Bernieri, 1990). Furthermore, in

genetically informed adoption and twin research, degree of biological relatedness between individuals, not specific molecular markers of genetically linked characteristics, is often the primary focus, and it has proven difficult to identify genes responsible for the heritability of complex traits, such as parenting (Plomin, 2012; Wahlsten, 2012)—although this situation is changing with respect to identifying specific genes (see Bakermans-Kranenburg & van IJzendoorn, 2008; Burkhouse, Gibb, Coles, Knopik, & McGeary, 2011; S. S. Lee et al., 2010; Mileva-Seitz et al., 2011; van IJzendoorn, Bakermans-Kranenburg, & Mesman, 2008). As to how genes might manifest in parenting, it is possible they do so through hormones implicated in childbearing and child rearing or because they contribute to the mental and emotional composition of the parent (both of these determinants of parenting are discussed next). However, contemporary parenting research is concerned with a broader range of questions than concern most heritability studies. Hormones The expression of parenting has been linked to hormones that are sometimes homologous in females and males (Bales, 2014; Corter & Fleming, 2002; Reburn & Wynne-Edwards, 1999). In most mammals, postpartum females respond to young differently than do nulliparous (virgin) females (Numan, 2012; Numan, Fleming, & Levy, 2006; Numan & Insel, 2003). Primiparous females are attracted to infants and care for them, whereas virgins typically avoid or reject infants, suggesting that hormonal events involved in parturition prime the brain to be sensitive to a new and unique set of stimuli (Lambert & Kinsley, 2012). Such hormonal changes enter the brain and act on neurons to modify brain function (Numan & Insel, 2003). A “female mammal's physiological state primes and activates neural and neurochemical mechanisms that promote maternal responsiveness [and]…dysregulation of such mechanisms would lead to inadequate maternal behavior” (Numan, 2012, p. 105). Perhaps the most studied hormone is oxytocin, which promotes prosocial behavior (Galbally, Lewis, van IJzendoorn, & Permezel, 2011) and is known to support the parent-infant bond in mammals (MacDonald & MacDonald, 2010). OT has been implicated in attachment and positive parenting in different species from rodent (Champagne, Diorio, Sharma, & Meaney, 2001) to monkey (Maestripieri, Hoffman, Anderson, Carter, & Higley, 2009) to human (Feldman, Gordon, & Zagoory-Sharon, 2011). Variation in the OT receptor gene is associated with maternal sensitivity (Bakermans-Kranenburg & Van IJzendoorn, 2008). Naber and colleagues (2010) videorecorded fathers during a play session with their toddlers following intranasal OT administration and then after intranasal placebo administration. Fathers increased their support for learning and exploration with respect for the child's autonomy following OT administration. Feldman et al. (2012) assayed OT from adults who were also genotyped for oxytocin receptor (OXTR) and CD38 risk alleles associated with social dysfunction. (CD38 is an ectoenzyme that mediates the release of brain OT.) Parent-infant interactions were microcoded for touch and gaze synchrony, and participants reported on parental care in childhood. Reduced plasma OT and both OXTR and CD38 risk alleles were related to less parental touch, and parents who reported less parental care showed lower plasma OT, low-risk CD38 alleles, and less touch, suggesting that peripheral and genetic

markers of the extended OT pathway might underpin core behaviors associated with human parenting. In one extension of this work, the same team observed associations between baseline OT levels in mothers and fathers and different parenting behaviors (engagement, affect synchrony, and communication). OT revealed differential associations with sexually dimorphic patterns of parental behavior: Mothers' baseline OT levels were positively correlated with affectionate but not stimulatory contact, whereas fathers' baseline OT levels were positively correlated with stimulatory but not affectionate contact (Feldman, Gordon, Schneiderman, Weisman, & Zagoory-Sharon, 2010; Feldman et al., 2011). Utilizing a doubleblind, placebo-controlled crossover design, Weisman, Zagoory-Sharon, and Feldman (2012) observed fathers and their 5-month-old infants twice in a face-to-face still-face paradigm following administration of OT or placebo to the father. OT administration increased fathers' salivary OT and key parenting behaviors that support parent-infant bonding. Moreover, parallel increases were found in infant salivary OT and engagement, including social gaze, exploration, and social reciprocity. In other words, OT administration had parallel effects on the treated parent and untreated child. Several other neurohormones show associations with parenting cognitions and practices. Decreased cortisol relates to high ratings on control and warmth, decreased emotionality in responding to children, and reduced conflict with children about rules. Mothers with higher cortisol levels are withdrawn and less interactive with children (Flinn et al., 2012). Fathers' cortisol levels are elevated at home on evenings that follow more stressful interactions with supervisors and with coworkers (Repetti, Robles, Reynolds, & Sears 2012; Saxbe, Repetti, & Nishina, 2008). Sprengelmeyer et al. (2009) implicated the likely involvement of estrogen and progestogen in mothers' emotional responses to baby faces. They suggested that neural substrates involved in reward processing and maternal behavior are possible candidate structures linked to these processes (Panksepp, 1998). Increases in prolactin (PRL) levels have been implicated in the expression of parenting in females as well as males, and fathers have higher PRL levels than nonfathers (Gettler, McDade, Feranil, & Kuzawa, 2012). Variation in the vasopressin receptor 1A (AVPR1A) is associated with maternal focused support (Avinun, Ebstein, & Knafo, 2012). Whereas estradiol plays a role in regulating maternal behaviors in female animals, the conversion of testosterone (T) into estradiol is believed to be involved in regulating paternal behavior in male animals (Trainor & Marler, 2001, 2002) and may constitute an endocrinological basis of paternal care (Fernandez-Duque, Valeggia, & Mendoza, 2009). Human fathers show different peripheral steroid hormone levels compared with childless men (Storey, Walsh, Quinton, & Wynne-Edwards, 2000). In men, T is elevated during mate acquisition (and conspecific competition) but downregulated in parenting (Gettler, McDade, & Kuzawa, 2011). A large representative study in the Philippines showed that, among single nonfathers, men with high waking T were more likely to become partnered fathers by the time of follow-up 4.5 years later. Men who became partnered fathers then experienced large declines in waking and evening T, declines that were significantly greater than declines in single nonfathers (Gettler, McDade, Feranil, & Kuzawa, 2011). Men with higher T levels report less sympathy in response to an infant cry (Fleming & Li, 2002), and fathers with higher

T engage in less caregiving (Alvergne, Faurie, & Raymond, 2009; Kuzawa, Gettler, Muller, McDade, & Feranil, 2009). In considering the possible implications of pregnancy-related hormones and other caregivinginvolved hormones, it is well to remember that many individuals who do not give birth to children—fathers, grandparents, in-laws, adoptive parents—not only parent them but come to love and develop strong emotional bonds with them too (Rilling, 2013). The Nervous Systems Just as genes and hormones are wrapped up in parenting, so are the structure and function of the autonomic and central nervous systems (Bridges, 2008; Brunton & Russell, 2008; Numan & Insel, 2003). Esposito and his colleagues (2014) measured autonomic physiological arousal using the novel technique of infrared thermography while Italian and Japanese adults viewed unfamiliar infant and adult faces of ingroup versus outgroup members. Both Italian and Japanese showed significant physiological activation (increase of facial temperature) for both ingroup and outgroup infant faces. Arousal responses to infants are mediated by the autonomic nervous system. These same investigators also used three convergent methodologies – Galvanic Skin Response (GSR), cardiac dynamics via Inter-Beat Interval (IBI), and right hand temperature change (RHTC) in fathers of typically developing children and non-fathers listening to the cries of infants later diagnosed with ASD compared to typical infant cries and laughter. They found that fathers and non-fathers alike show greater negative responses (increased GSR) to ASD compared to typical infant cries and laughter. Fathers showed higher IBI and greater RHTC than non-fathers while listening to typical and atypical cries. Thus, fathers and non-fathers show emotional arousal mediated by sympathetic activation while listening to cries of children with ASD (Esposito et al., 2015). Parenthood is known to alter the adult brain (Barrett & Fleming, 2011; Bornstein, 2013a; P. Kim et al., 2010a; Leuner, Glasper, & Gould, 2010; Rilling, 2013). For example, Bornstein, Arterberry, and Mash (2013) found that just 3 months of exposure to their own infant's face changes the visual cortex evoked response potential (ERP) in new mothers. Nonhuman animal studies show that the mammalian forebrain plays an important part in the expression of reproductive behavior and specific behaviors during nursing and care of offspring, ones that are different in female and male animals and are reflected in functional and morphologic sex differences in neural forebrain circuitry (Champagne et al., 2001; Corter & Fleming, 2002; Sheehan & Numan, 2002; Simerly, 2002). The physiological state associated with parturition causes a functional rerouting of the brain circuits over which infant stimuli are processed. In virgins, infant stimuli have a negative valence and activate neural circuits that regulate avoidance and defensiveness, resulting in antisocial responses. Such mechanisms prevent the virgin from caring for young when she is not lactating. In postpartum females, infant stimuli have a positive valence and activate neural circuits that regulate attraction, approach, and acceptance behaviors (prosocial maternal responses). In humans, the neuroscience of parenting is a cresting area of research interest (Bornstein, 2013a). Using visual (usually face) and auditory (usually cry or laugh) stimuli (of own and

unfamiliar infants and children), imaging studies from electroencephalography to ERP to magnetoencephalography to functional magnetic resonance imaging have begun to expose neural activation patterns in specific regions of the brain that are associated with select parenting cognitions and practices leading to incipient models of a complex brain network hypothesized to mediate human parenting (Swain, Lorberbaum, Kose, & Strathearn, 2007). Since Bartels and Zeki (2000, 2004), programmatic research with own versus other baby photographs and videos (Bornstein et al., 2013; Leibenluft, Gobbini, Harrison, & Haxby, 2004; Noriuchi, Kikuchi, & Senoo, 2008; Ranote et al., 2004; Strathearn, Li, Fonagy, & Montague, 2008; Swain, Leckman, Mayes, Feldman, & Schultz, 2006) and with cries of own infant versus standard cries versus control noises (De Pisapia et al., 2013; Ranote et al., 2004; Venuti et al., 2012) has revealed enhanced activations in regions of mothers', fathers', and nonparents' brains associated with empathy, responsiveness, and emotion recognition and evaluation. Human mothers show neural activation in limbic forebrain structures in response to infant crying compared with neutral sounds (Lorberbaum et al., 1999) and bilateral activation of the orbitofrontal cortex while viewing pictures of their own versus unfamiliar infants (Nitschke et al., 2004). First-time mothers activate mirror-neuron regions of the brain more strongly when observing and empathizing with their own compared with unknown infants (Lenzi et al., 2009). Connecting brain activation with cognitions, mothers viewing smiling pictures of their own versus unfamiliar infants yields brain activation that is correlated with pleasant mood ratings and affective responses to their infant (Nitschke et al., 2004). When Laurent and Ablow (2012) investigated depression-related differences in primiparous mothers' neural responses to their own infant's distress cues using functional neuroimaging, they found that nondepressed mothers activated to their own infant's cry greater than to a control sound in a distributed network of paralimbic and prefrontal regions, whereas mothers diagnosed with major depressive disorder failed to show activation. Depressed compared to nondepressed mothers also showed less striatal (caudate, nucleus accumbens) and medial thalamic activation. Experience-independent and experience-dependent neuroplastic alignments of the human brain likely subserve the biological requirements of child care, securing offspring survival and well-being and perhaps fostering reproductive fitness in subsequent generations (Bornstein, 2013a; P. Kim et al., 2010b; Rilling, 2013). Papoušek and Papoušek (2002) advanced the notion that some parenting is likely hardwired in human beings. Intuitive parenting involves responses that are developmentally suited to the age and abilities of the child and that likely have the goal of enhancing child adaptation and development. Parents regularly enact intuitive parenting programs in an unconscious fashion; such programs do not require the time and effort typical of conscious decision making; and, being more rapid and efficient, they utilize less attentional reserve. An example of such intuitive parenting is child-directed speech (Soderstrom, 2007) whose special characteristics vary from adult-directed speech along prosodic, simplicity, redundancy, lexical, and content dimensions. Cross-cultural study attests that child-directed speech is (essentially) universal (Jacobson, Boersma, Fields, & Olson, 1983; Snow, 1977; but see Ratner & Pye, 1984). When communicating with their children, even deaf mothers modify their sign language the way hearing mothers use child-directed speech (Erting, Prezioso, & Hynes, 1994). Indeed, parents find it difficult to resist or modify such intuitive behaviors, even when asked to do so

(Trevarthen, 1979). Additional support for the premise that such interactions with children are intuitive comes from observations that nonparents (males and females) who have little prior experience with children modify their speech as parents do when in the presence of a young child and even when asked to imagine speaking to one (Jacobson et al., 1983). Many parenting cognitions and practices are likewise unconscious, habitual, and possibly thoughtless, and they are enacted by parents and nonparents alike (Senese et al., 2013).

Psychological Characteristics In addition to these (and possibly other) biological characteristics, a variety of psychological characteristics in the parent has been identified that shape parenting. They include the gender, age and stage, health status, cognition (including parenting cognitions), personality, intergenerational transmission and family of origin, and experiences of the parent. Gender Parent gender is a perennially important topic (Bornstein, 2013b), especially as the role of fathers has become more widely recognized (Ribas & Bornstein, 2005). Some salient distinctions between mothering and fathering already have been drawn. Mothers and fathers hold some similar and some different ideas about parenting and appear to interact with and care for children in some convergent and some complementary ways; that is, they tend to divide some of the labor of caregiving and engage children by modeling different orientations and emphasizing different types of interactions. Mothers are more often involved in direct caregiving and typically provide children with more close emotional support, whereas fathers often provide more informational and material support and involve children more in leisure activities (Collins & Russell, 1991; Parke & Buriel, 2006; Steinberg & Silk, 2002). Mothers tend to be more knowledgeable about children's friendships and their social network than fathers; fathers are more likely to be perceived as relatively distant authority figures who may be consulted for objective information (such as help with homework) but who are less frequently sought out for relationship support or guidance (such as help for problems with friends) than mothers (Updegraff, McHale, Crouter, & Kupanoff, 2001; Waizenhofer, Buchanan, & Jackson-Newsom, 2004). Boivin and colleagues (2005) analyzed the factor structure of a self-administered questionnaire designed to assess 2,122 mothers' and 1,829 fathers' self-efficacy, perceived impact, hostile-reactive behaviors, and overprotection vis-àvis their young infants. The two parents did not differ with respect to perceived parental impact, but these researchers found significant gender differences in other parenting selfperceptions. Mothers were more worried about the health and safety of their infants and felt more effective as parents than did fathers, and fathers were more prone to hostile-reactive behaviors than mothers. Across nine countries, mothers report more progressive parenting attitudes than fathers, and fathers report more authoritarian attitudes than mothers (Bornstein, Putnick, & Lansford, 2011). However, mothers and fathers may differ in their consistency, favoring mothers (Belsky, Taylor, & Rovine, 1984; Lytton & Zwirner, 1975). Anticipating evolutionary considerations discussed toward the conclusion of this chapter, the bias toward caregiving in women could be evolutionally advantageous, considering that

women are primary caregivers of children in most societies (Eibl-Eibesfeldt, 1989). Evolutionary psychology attributes these consistent differences in mothering and fathering among mammals principally to maternal internal gestation and obligatory postpartum suckling (Clutton-Brock, 1989). Differences in the relative costs and benefits of producing offspring are argued to play key parts in understanding the evolution of gender differences in reproductive strategies and in parental investment (Trivers, 1972): The gender with the lower potential rate of reproduction invests more in parenting efforts (Clutton-Brock & Vincent, 1991). Age and Stage of Life The average age at first birth in the United States is 26.0 years. The contemporary demographics of parturition in the United States indicate that the rate of teenage (15–19 years) motherhood is epidemic (273,105 live births in the United States in 2013; CDC/National Center for Health Statistics, 2015), as approximately 1 in 3 adolescent women becomes pregnant by the end of her nineteenth year. At the same time, increasing numbers of adult women are delaying conception, extending the age range for pregnancy and birth (J. A. Martin et al., 2013). In 2011, for the first time on record, for example, birth rates in Canada were higher for women in their late 30s than in their early 20s. These demographic trends might be ascribable to several factors that contribute to the decision to bear a child and so to parent (Golombok, 2002, 2013), and they apply to first-time parenthood. Parents also age as their children do, and they encounter their own life challenges along the parentway. For example, the typical parent may be 40 years of age or older when her or his first child enters adolescence, and parents at midlife often encounter unique developmental needs brought on by their own passage into midlife at the same time as their children (Bornstein, Jager, & Steinberg, 2012). This period in the family life cycle sometimes proves to be a low point in parents' marital and life satisfaction as well as a period of heightened risk for divorce (Gecas & Seff, 1990; Gottman & Levenson, 2000). These demographic trends, in turn, raise questions about specific associations that may obtain between age or stage of life and parenting. It was once standard to believe that optimal childbearing takes place between about 20 and about 30 years of age (Rindfuss & Bumpass, 1978) and support for this ∩-shaped function still prevails (McGrath et al., 2014). Rossi (1980) proposed a “timing-of-events” model that suggests that having a child when very young or very old might represent “off-time” versus “on-time” variations in childbearing that result in decreased social reinforcement and lead to curvilinear relations between age and parenting. For example, teenage and older mothers tend to report higher levels of parenting stress compared to mothers in their 20s and early 30s (Östberg & Hagekull, 2000). Whether and how parent chronological age relates to parenting cognitions or practices appears to depend on specifics of the assessment. For some parenting variables, age does not seem to matter. Mothers of all ages possess implicit beliefs (Holden & Buck, 2002), and mothers of all ages engage equally in child-directed speech (Papoušek & Bornstein, 1992). For many other forms of parenting, however, age matters a lot. The biopsychosocial impacts of early childbirth are fairly well established. Adolescent mothers experience more pregnancy and delivery problems, bear less healthy babies, express less desirable or realistic child-rearing cognitions,

hold less realistic expectations about child development, and parent using less favorable practices than do adult mothers (Berlin, Brady-Smith, & Brooks-Gunn, 2002; Bornstein & Putnick, 2007; Bornstein, Putnick, Suwalsky, & Gini, 2006; Coley & Chase-Lansdale, 1998; Demers, Bernier, Tarabulsy, & Provost, 2010; Moore & Brooks-Gunn, 2002; Pomerleau, Scuccimarri, & Malcuit, 2003; Schlomer & Belsky, 2012). For example, in a multivariate analyses that compared adolescent (≤19 years) and adult (≥26 years) mothers, controlling for sociodemographic and other human and social capital factors, Y. Lee (2009) found that adolescent mothers were more psychologically and physically (spanking) aggressive. Lewin, Mitchell, and Ronzio (2013) used the nationally representative Early Childhood Longitudinal Study-Birth cohort data set to compare parenting practices of adolescent mothers (25 years) when their children were 2 years of age. Controlling for socioeconomic differences, adolescent mothers exhibited less supportiveness, sensitivity, and positive regard than emerging adult mothers, who exhibited less than adults. Adolescent and emerging-adult mothers reported equivalent frequencies of spanking and use of timeout, significantly more than adult mothers. Some effects of aging on parenting are thus clear cut, but others not. A 35-year-old woman has a 1 in 400 chance of conceiving a child with Down syndrome, and this likelihood increases to 1 in 110 by age 40 and then to approximately 1 in 35 by the age of 45 (National Down Syndrome Society, http://www.ndss.org/). Tending and rearing children are physically challenging, and older primiparas may possess relatively diminished physical capacities to meet those demands (Mirowsky, 2002). However, older mothers are more likely to adhere to good diets and gain weight appropriately during pregnancy, to begin prenatal care in the first trimester of pregnancy, and to avoid toxins (smoking and alcohol) during pregnancy. Age is often conceived of as a marker for maturity, perspective, and patience; older-adult mothers tend to possess more experience and information and may feel more psychologically ready to assume the many responsibilities of child rearing. In the view of some, mothers' readiness to be a parent is determined by competency and maturity as measured by her age at childs' birth (Garrett, Ferron, Ng'andu, Bryant, & Harbin, 1994). The more mature and experienced mothers are, the more appropriate and propitious (it is commonly believed) their parenting beliefs and behaviors are likely to be (Bornstein, Putnick, & Suwalsky 2012; Demers, Bernier, Tarabulsy, & Provost, 2010). Even after controlling for SES, maternal education, ethnicity, intelligence, social desirability of responding, and child age and language competence, older mothers direct more praise and physical affection toward their toddlers than do younger mothers (Bornstein & Putnick, 2007). A study of European American mother-child dyads found that older mothers were more sensitive and optimally structuring with their infants and toddlers than were younger mothers (Bornstein, Putnick, & Suwalsky, 2012). Fathers' age matters as well in these and other ways. A monotonic association obtains between advancing paternal age and risk of autism spectrum disorders (ASD) in children (Reichenberg et al., 2006), for example, and a study of genome-wide mutation rates by sequencing the entire genomes of a sample of Icelandic parent-offspring trios revealed that the diversity in mutation rate of single nucleotide polymorphisms is dominated by the age of the father at a child's conception, implicating increases of about two mutations per year (Kong et al., 2012).

Advancing paternal age is specifically associated with increased genetic mutations during spermatogenesis, which may cause psychiatric morbidity in offspring. A population-based cohort study of all people born in Sweden from 1973 to 2001 (N = 2,615,081) revealed that paternal age is associated with increased risk of some psychiatric disorders (ASD, psychosis, and bipolar disorders) but decreased risk of the other indexes of morbidity. In another withinsibling comparison, offspring of fathers 45 years and older were at heightened risk of ASD, ADHD, psychosis, bipolar disorder, substance use problems, failing a grade, and low educational attainment compared with offspring of fathers 20 to 24 years old (D'Onofrio et al., 2014). Of course, age is a “social address” (Bronfenbrenner & Crouter, 1983), and age usually stands only as a proxy for underlying processes. More proximal intrapersonal factors likely play the meaningful roles in parenting, and it is possible to identify some critical developmental phenomena that would help to mark mature caregiving. For example, developing executive functions (discussed later) coordinate metacognitive processes through monitoring and controlling the use of knowledge and strategies that redound positively to parenting (Barkley, 2012; Yeager & Yeager, 2013). Health Status To parent well, a parent's own needs must be met. When women are inadequately nourished, for example, their health and social development may be compromised, and their abilities to bear and rear healthy children are jeopardized (Houck, Rodrigue, & Lobato, 2007). Malnourished women fall ill more often, and they have smaller babies, whereas women whose diets are rich in protein have fewer complications during pregnancy, transit shorter labors, and bear healthier babies. Deficiencies in zinc, folic acid, and protein have been linked to central nervous system dysfunction, prematurity, and low birthweight (Keen, Bendich, & Willhite, 1993). In many regions of the developing world, malnutrition is a chronic problem (UNICEF, 2006), but this problem is not limited to poor countries. The U.S. Special Supplemental Food Program for Women, Infants, and Children (WIC) provides food packages, nutrition education, and health care to low-income pregnant women and new mothers. More than 8 million people receive WIC benefits each month (Lino, 2013). Severe malnutrition in early pregnancy (especially lack of folic acid, a B vitamin) increases the risk of neural tube defects, such as spina bifida, in which the spinal cord does not close completely, or anencephaly, in which part of the brain does not develop. Malnutrition later in pregnancy is associated with low birthweight. But the long-term prognosis for children chronically malnourished in utero is not entirely bleak. If children receive nutritious diets before age 2, they can rebound. Near the end of World War II, western (but not northern or eastern) Holland endured a food blockade that provided an unfortunate but important natural experiment in maternal health. Unlike other famines, the so-called Dutch Hunger Winter struck during a precisely circumscribed time and place and in a society that kept comprehensive and meticulous health records. As a result, researchers later identified children who were exposed to fetal malnutrition in different trimesters and could follow them through adolescence and well into adulthood. In general, children born during this time of famine did not suffer pervasive long-term physical or mental

disabilities (Hoek, Brown, & Susser, 1999; Susser, Hoek, & Brown, 1998). Relative to those who received proper nutrition, however, malnourished children suffered more nervous system congenital abnormalities as well as increased risk of schizophrenia. Health effects are not restricted to mothers. Diet folate deficiency is associated with giving birth to a child with a birth defect. Male mice that were fed a diet containing less than 15% of the recommended amount of folate for their lifetime (from the time they were fetuses through to their reproductive maturity) showed deficiencies in fertility as well as litter offspring with approximately 30% more birth defects than controls (Lambrot et al., 2013). Cognition Parents' attentiveness, intelligence, mental functioning, and (as we see later) even memories of their own childhood help to create a “strategy frame” or “affective lens” that influences parenting (Putallaz, Costanzo, & Smith, 1991). Attention at neurobiological levels is heightened in new mothers (Purhonen, Valkonen-Korhonen, & Lehtonen, 2008); policewomen report enhanced vigilance following the birth of their first child (Fullgrabe, 2002); and women's improved attentional processing of infant emotions during pregnancy relates to their later relationships with their infant (Pearson, Lightman, & Evans, 2011). Analyzing data from the Massachusetts site of the National Institute of Child Health and Human Development (NICHD) Study of Early Child Care and Youth Development, Mulvaney, McCartney, Bub, and Marshall (2006) learned that mothers' verbal intelligence predicted the effectiveness of their scaffolding collaborations with children (which in turn uniquely predicted later cognitive capabilities of the children). Executive functions include self-regulation, sequencing, flexibility, response and inhibition, planning, and organization of behavior (Butterfield, Albertson, & Johnston, 1995; Denckla & Reiss, 1997; Eslinger, 1996). The orderly approach to problems, maintenance of problem solving sets for future goals, control processes for organizing behavior over time, flexibility and effectiveness of verbal self-regulation, skillful use of strategy, and behaviors that alter the likelihood of later events are all executive functions…and describe well the requirements of parenting. Grattan and Eslinger (1992) suggested that mental flexibility and perspective-taking are cognitive prerequisites to empathic understanding; impulse control, temporal integration, and synthesis of multiple pieces of information are cognitive prerequisites to identity formation; and symbolic thinking, weighing alternative possibilities, and considering consequences among alternatives are cognitive prerequisites to moral maturity. Thus, executive functions exert powerful and pervasive influences on social behavior, and immaturity or impairment of executive function can lead to demanding and self-centered behavior, lack of social tact and restraint, impulsive speech and actions, disinhibition, apathy and indifference, and absence of empathy, all of which are hallmarks of dysregulated parenting. The neuropsychological underpinnings of executive functions are usually assigned to the prefrontal cortex and its extended networks (Pennington, Bennetto, McAleer, & Roberts, 1996). Notably, individuals with localized prefrontal injury display poor parenting (Eslinger, Grattan, Damasio, & Damasio, 1992). Consider patient DT, who:

proved unable to anticipate and meet her child's needs, such as planning meals, changing clothing, and providing nurturance and comfort… Her performance [was] erratic, impulsive, and marked by poor follow through on required tasks, failure to learn from mistakes, and very negative reactions to criticism… She [had] very limited capacity for empathic understanding, inadequate identity development, difficulties in vocational adjustment, and a concrete level of moral reasoning. (Grattan & Eslinger, 1992, p. 185) DT shows how the want of executive functions profoundly undermines parenting in an individual who otherwise possesses normal motor and sensory functions and broadly normal intellect, perception, language, and memory. To this point, it is important to note that the prefrontal cortex shows a prolonged course of development, with changes in synaptic density detectable even into the mid-20s (Shaw et al., 2006), and so could very well underlie the wellknown challenges commonly associated with adolescent parenting. As indicated earlier, teenage mothers display notable parenting deficits (Loeber, Burke, Lahey, Winters, & Zera, 2000). Younger mothers score lower than older mothers on a construct called cognitive readiness to parent (aggregated knowledge of child development, parenting attitudes, and parenting style), and cognitive readiness to parent predicts parenting, at least during infancy (Whitman, Borkowski, Keogh, & Weed, 2001), and accounts for relations between early maternal interactions and 1-year attachment status (Lounds, Borkowski, Whitman, Maxwell, & Weed, 2005). Parenting Cognitions As reviewed earlier, parenting cognitions come in many flavors and are generally believed to serve many functions: They affect parents' sense of self, help to organize parenting, and mediate the effectiveness of parenting. Cognitions help to structure parents' practices and determine how much time, effort, and energy to expend in parenting. To begin at the beginning, however, cognitions instigate whether to become a parent. Because of the widespread availability of contraception and assisted reproductive technologies, decisions to bear and rear children are elective. Changing cultural mores also permit (15%–19% and increasing) people to remain voluntarily childless (Letherby, 2002; Livingston & Cohn, 2010). Today, the rates of unintended and intentional pregnancies are about the same ( 50% of each in the United States; Finer & Zolna, 2011). Notably, unplanned pregnancies are associated with many poorer forms of parenting (Treyvaud, Rogers, Matthews, & Allen, 2010), including increased likelihood of perinatal mood problems and reduced well-being in the postpartum period (Grussu, Quatraro, & Nasta, 2005; Lachance-Grzela & Bouchard, 2009; Yanikkerem, Ay, & Piro, 2013), less time and attention paid to infants (Barber, Axinn, & Thornton, 1999), and more negative parenting (East, Chien, & Barber, 2012). How parents see themselves vis-à-vis children generally can lead to their expressing one or another kind of affect, thinking, or behavior in child rearing. According to the Civitas Initiative et al. (2000) survey, for example, 90% of new parents in the United States have confidence in their abilities and think of themselves generally as good parents. Parents with a low perceived balance of power within the caregiving relationship may see unresponsive children as behaving with intent, an interpretation that easily fosters in parents a sense of helplessness and

ultimately anger (Bugental & Happaney, 2004). Especially salient self-perceptions of parents have to do with their feelings of competence in the role of caregiver, satisfaction gained from caregiving, investment in caregiving, and balance of caregiving with other social roles. Most is known about parenting competence (Bornstein, Haynes, et al., 2004; Bornstein, Hendricks, et al., 2003). Among low-income families, mothers with lower levels of belief in parents as teachers are less likely to elaborate appropriately with their children during shared book reading (DeBaryshe, 1995). Functionally, perceptions of competence in parenting are associated with parents' use of effective parenting strategies (Johnston & Mash, 1989; Teti & Gelfand, 1991). Self-efficacy theory posits that adults who evaluate themselves as competent, who know what they can do, and who understand the likely effects of their actions will, as parents, more likely act as constructive partners in their children's development (Bandura, 1986, 1989; Coleman & Karraker, 1998; Conrad, Gross, Fogg, & Ruchala, 1992; King & Elder, 1998). Eccles and Harold (1996) reported that parents' confidence in their ability to influence their children's academic performance and school achievement is associated with parents' school involvement and predicts parents' helping with their children's academic interests (Epstein & Sanders, 2002; see, too, Hoover-Dempsey & Sandler, 1997). Mothers who consider themselves efficacious and competent in their role as parents tend to be more responsive and empathic and less punitive and hold more appropriate developmental expectations (De Haan, Prinzie, & Deković, 2009; Meunier, Roskam, & Browne, 2011). Parents who believe that they can or cannot affect characteristics in children likely act accordingly. The Civitas Initiative et al. (2000) reported that approximately 25% of parents in the United States thinks that a baby is born with a certain level of intelligence that cannot be increased or decreased by how parents interact with the baby. How parents see their own children has specific consequences as well, of course. Mothers who regard their child as being difficult, for example, are less likely to pay attention or respond to their child's overtures (Putnam, Sanson, & Rothbart, 2002). Other prominent parenting cognitions likewise influence parenting. Attitudes refer to an individual's predispositions, reactions to, or affective evaluations of the supposed facts about an object or situation. Expectations about a child as well as developmental norms and milestones—when a child should achieve a particular skill—affect parents' appraisals of their child and their parenting. Expectations also generally bias observations so that information consistent with the expectancy is more likely to be attended to, processed, and acted upon. Jamaican mothers expect their children to sit and to walk early, whereas Indian mothers living in the same city expect their children to crawl later. In each case, children's actual attainments of developmental milestones accord with their mothers' expectations (Hopkins & Westra, 1989, 1990). Attributions are interpretations of causations of events and behaviors, and parenting attributions refer to assigned meanings and definitions of a child's behavior and so can affect parents' caregiving practices (Bugental & Happaney, 2002). Parental attributions typically distinguish between the internal (intentional) and the external (situational). An internal attribution might refer to parental interpretations of a child's behavior as dispositional and deliberate, whereas an external attribution might refer to parental interpretations of a child's behavior as contextual, transitory, and even accidental (Coplan, Hastings, LagacéSéguin, & Moulton, 2002). In certain circumstances, parents might believe a child is behaving

purposefully, when the child's behavior may in fact be developmentally typical. For example, higher levels of internal attributions of child misbehaviors are more prevalent among neglectful, abusive, and authoritarian mothers (Z. Wang, Deater-Deckard, & Bell, 2013). Investment in, involvement with, and commitment to children is foundational to positive child rearing; indeed, Baumrind and Thompson (2002) defined ethical parenting “above all [as] requiring of parents enduring investment and commitment throughout their children's long period of dependency” (p. 3). Invested parents are more responsive and view their children more positively (Greenberger & Goldberg, 1989). In turn, parental investment in children's lives and parental responsibility for children's care help to ensure that parents provide children with medical attention, physical activity, and proper nutrition they require for wholesome development (Cox & Harter, 2003). Finally, how parents balance their multiple roles in life—parent, spouse, and employee—reflects on their effectiveness in those diverse roles, including their parenting (Perry-Jenkins, Repetti, & Crouter, 2000). Mothers who are happy with their roles are more accepting of their children (Stuckey, McGhee, & Bell, 1982). Parenting satisfaction affords a sense of well-being in the parenting role that translates to positive emotionally available parent-child interactions (Bornstein, Putnick, et al., 2012). Whereas goals, attitudes, and other similar cognitions may or may not be factual, parenting knowledge draws on the science base as well as social construction and is thought to be valid and reliable by members of the clinical and research communities (Bornstein et al., 2010). Parenting knowledge of child rearing and child development encompasses many domains, including fulfilling the biological and physical as well as socioemotional and cognitive needs of children as they develop, understanding of normative child development, and awareness of practices and strategies for maintaining and promoting children's health and coping effectively with children's illness (Bornstein & Cote, 2004b; Scrimin et al., 2009). The general state of parents' knowledge affects their everyday decisions about their children's care and upbringing and guides parents' decisions about how to maintain child health and when and how to seek care if children fall ill (Hickson & Clayton, 2002; Melamed, 2002; Zuckerman & Keder, 2015). Parenting knowledge equips parents with information they need to interpret children's abilities and accomplishments and to tailor their interactions accordingly. In turn, parenting knowledge is associated with enhanced parental self-perceptions of competence, satisfaction, and investment in parenting (Bornstein, Hendricks, et al., 2003). For example, the relation between parental self-efficacy and parenting competence is moderated by parenting knowledge: Parental self-efficacy and parenting competence are positively associated when parenting knowledge is high; by contrast, self-efficacy and competence are inversely associated when knowledge is low (Hess, Teti, & Hussey-Gardner, 2004). Parents who harbor unrealistic developmental expectations that are not informed by accurate knowledge experience greater stress as a result of mismatches between expectations and actual child behaviors (Teti & Gelfand, 1991), whereas knowledgeable parents have more realistic expectations and are more likely to behave in developmentally appropriate ways with their children (Grusec & Goodnow, 1994). Mothers' specific knowledge of child rearing and child development explains variations in their emotional relationships with children (Bornstein, Putnick, et al., 2012). Safety is measured in terms of awareness of health care, the identification and treatment of illnesses, and accident prevention, and these beliefs guide

parents' decisions about how to maintain child health. Using the prospective longitudinal study data set of a low-birthweight preterm cohort from the multisite Infant Health and Development Program, Benasich and Brooks-Gunn (1996) found that maternal knowledge of child development and child rearing conditioned the quality and structure of the home environment mothers provided (which in turn affected child cognitive and behavioral outcomes). Sudden infant death syndrome (SIDS) is the leading cause of reported neonatal mortality in the United States (http://www.cdc.gov/sids/), and sleeping prone or on too-soft bedding that may cover an infant's mouth and nose increases the likelihood of asphyxiation (Scheers, Rutherford, & Kemp, 2003). Still, approximately 20% of parents place infants ages 1 to 3 months on their stomachs during sleep (Gibson, Dembofsky, Rubin, & Greenspan, 2000). In all these different ways, parents' cognitions can lead to different transactions with children. In what may be called the standard model of parenting (widely assumed but seldom appropriately tested), parents' cognitions are hypothesized to prompt or direct parents' practices and, ultimately, children's development. That said, relations between beliefs and behaviors are historically an unsettled area in social psychology (Armitage & Conner, 2001; Festinger, 1964; Green, 1954; LaPiere, 1934; Wicker, 1969), and relations between parental beliefs and behaviors specifically are equally elusive (Bornstein, Cote, & Venuti, 2001; Coleman & Karraker, 2003; Cote & Bornstein, 2000; Holden, 2002; S. G. Miller, 1988; Okagaki & Bingham, 2005; Sigel & McGillicuddy-De Lisi, 2002). Many researchers have reported no relations between mothers' professed parenting attitudes and their activities with their children (Cote & Bornstein, 2000; McGillicuddy-De Lisi, 1992), where others have reported associations between parents' beliefs and behaviors that are relatively weak or positive but nonsignificant (Coleman & Karraker, 2003; Mantzicopoulos, 1997; Sigel & McGillicuddy-De Lisi, 2002). Those parenting beliefs and behaviors whose consonance has been evaluated have often been very general, giving little reason to expect an association. When more circumscribed and conceptually corresponding cognition-practice domains are studied, however, some maternal child-rearing beliefs have been found to relate to some self-reported or observed child-rearing behaviors, supporting expected links in the putative standard model causal chain (Coplan et al., 2002; K. Y. Huang, O'Brien Caughy, Genevro, & Miller, 2005; Kinlaw, Kurtz-Costes, & Goldman-Fraser, 2001; Strassberg & Treboux, 2000). For example, the Benasich and BrooksGunn (1996) study just reported found that maternal knowledge of child development and child rearing conditioned the quality and structure of the home environment mothers provided, which in turn affected child cognitive and behavioral outcomes. The strength of the association between parents' cognitions and practices appears to depend, therefore (at least in part) on alignment of conceptual links between the respective contents of their beliefs and their behaviors, as for example between authoritarian attitudes and discipline strategies and between intuitions about parenting effectiveness and competence. European American mothers endorse the importance of independence and self-assertiveness when asked to describe their ideal child, whereas Latina mothers accentuate obedience and respect for theirs (Harwood, Leyendecker, Carlson, Asencio, & Miller, 2002; Tamis-LeMonda & McFadden, 2010). In accord with these expressed values, in real-time interactions, U.S. mothers encourage their

children to feed themselves as early as 8 months of age, whereas Latina mothers hold their children close on their lap and control feeding from start to finish. In brief, beliefs do not always map to behaviors directly, but the two coexist in complex ways, and meaning assigned to each is critical (Bornstein, 1995). Parenting beliefs provide parents with a framework for interpreting their children's behaviors, guiding parents' interactions with their children and determining the activities and opportunities that parents supply. Ethnographic interviews of mothers with infants between the ages of 2 and 18 months revealed that some mothers avoid using physical punishment with infants because they believe that infants are not able to clearly understand right and wrong, whereas other mothers believe that infants can misbehave intentionally and need to be punished to stop their bad behavior and learn to respect the mother's authority. Subsequent quantitative analyses revealed that mothers who expressed concerns about bad behavior and spoiling interacted less positively with their infants (Burchinal, Skinner, & Reznick, 2010). It is especially noteworthy how early cognitions may function in the development of parenting. Siddiqui and Hägglöf (2000) found that mothers' antenatal attachment expectancy toward their unborn child predicted mothers' sensitivity at 3 months postpartum (see also Thun-Hohenstein, Wienerroither, Schreuer, Seim, & Wienerroither, 2008). Similarly, Haltigan and colleagues (2014) determined that mothers' attachment representations assessed prenatally predicted observed maternal sensitivity at 6 months postnatally. Using an equivalent longitudinal design, Dayton, Levendosky, Davidson, and Bogat (2010) learned that mothers who professed affectively disengaged prenatal representations of their children were at 1 year more behaviorally controlling; mothers whose representations were affectively distorted were hostile; and mothers with balanced representations demonstrated more positive parenting. Personality Two general orientations have guided theory and research linking personality to parenting: One concerns a factor structure thought to compose normal personality, and the other concerns more specific personality characteristics, from anxiety and stress to depression and psychopathology. Sigmund Freud (1949), the father of psychoanalytic theory, speculated that a parent's personality would determine the nature of parenting as well as the parent-child relationship and the child's development. Another consistent theme among psychoanalytic theorists is that, if parents' emotional needs are not met during the course of their own development, then their own psychological makeup would reflect in their parenting (Cohler & Paul, 2002; Holden & Buck, 2002). Anna Freud (1955/1970) described mothers who rejected their children, sometimes due to psychosis but more often because of their own neurotic conflicts. Personality therefore has a significant part to play in parenting: “One cannot take the ‘person’ out of the parent” (Vondra, Sysko, & Belsky, 2005, p. 2). Contemporary views derived from personality psychology assert that some parenting cognitions and practices reflect general and stable personality characteristics (Belsky & Barends, 2002). Bronfenbrenner and Morris (2006) opined that personality factors constitute person “force characteristics” likely to

influence children's development because personality affects parenting directly and because it colors other social contextual factors that influence parenting, including spouse selection and marital relationships. Personality is currently conceptualized as a profile of five broad-band factors (the so-called Big Five), each with lower-level facets (McAdams & Pals, 2006), that may be “universal” (Allik & McCrae, 2004; McCrae et al., 2005). Openness to Experience reflects a tendency to have a broad perspective and to approach life in intelligent, creative, philosophical, and inquisitive ways. Mothers who are more open are also more emotionally expressive with their children (C. L. Smith et al., 2007). Neuroticism reflects a proneness to psychological distress, unrealistic ideas, excessive cravings or urges, maladaptive coping responses, and a perturbable, insecure, and vulnerable orientation to life. Mothers who are neurotic display high levels of negative affectivity toward their children, are less responsive to their children's needs, engage in more physical and verbal power assertion, and are less encouraging of their children's autonomy (Clark, Kochanska, & Ready, 2000). Kotov, Gamez, Schmidt, and Watson (2010) reported that Neuroticism shows associations with internalizing and externalizing psychopathology (see also Decuyper, De Pauw, De Fruyt, De Bolle, & De Clereq, 2009; Ruiz, Pincus, & Schinka, 2008). Likewise, more neurotic fathers are more negative with their children, whereas less neurotic fathers seem to be less behaviorally and emotionally involved with theirs (S. Wang, Repetti, & Campos, 2011). Extraversion reflects the quantity and intensity of interpersonal interaction, activity level, need for stimulation, and capacity for joy, control, and assertiveness. Extraverted mothers are more nurturant and supportive of their children's autonomy (Clark et al., 2000; Losoya, Callor, Rowe, & Goldsmith, 1997; C. L. Smith et al., 2007) but also more controlling (i.e., power assertive and forceful in disciplinary style) than those who are less extraverted (Clark et al., 2000; Kochanska, Clark, & Goldman, 1997). Agreeableness reflects an interpersonal orientation in feelings, thoughts, and actions along a continuum from compassion to antagonism, the high end of which is characterized as cooperative, trusting, and warm. Agreeable mothers express higher levels of positive affect, support, and warmth with their children, are more sensitive and responsive to their children's needs, and are less likely to be affectively negative, uninvolved, and overreactive than less agreeable mothers (Belsky, Crnic, & Woodworth, 1995; Browne, Meunier, O'Connor, & Jenkins, 2012; de Haan, Prinzie, & Deković, 2009; Kochanska et al., 1997; C. L. Smith et al., 2007). Finally, Conscientiousness reflects the extent to which a person is well organized, responsible, decisive, dependable, hardworking, and even ambitious. Conscientious mothers provide more structure and use less forceful disciplinary styles than less conscientious mothers; however, conscientiousness is also associated with restrictive and overcontrolling child rearing (Clark et al., 2000; Neitzel & Stright, 2004; Verhoeven et al., 2007). A burgeoning literature explores relations between one or more of the Big Five and parenting (e.g., Koenig, Barry, & Kochanska, 2010), and a critical mass has accumulated to allow Prinzie, Stams, Dekovic, Reijntjes, and Belsky (2009) to conduct a meta-analysis that examined associations between parental personality and parenting practices. Results revealed only modest associations between parental personality and parenting warmth, behavioral control, and autonomy support. Additionally, this meta-analysis showed that several

moderators (e.g., parent age) explained variability among effect sizes reported across individual studies. Only a small number of studies has assessed all five personality factors in relation to parenting cognitions and/or parenting practices, controlling for one another. Factor analysis of a personality inventory completed by a community sample of European American mothers of firstborn 20-month-olds who also responded to multiple measures of parenting cognitions (knowledge, self-perceptions, and reports about behavior) and observed in interaction with their children for diverse parenting practices (language, sensitivity, affection, and play) replicated extraction of the Big Five personality factors and showed, controlling for sociodemographic characteristics, that the five personality factors (as variables and in patterns as clusters) related differently to diverse parenting cognitions and practices (Bornstein, Hahn, & Haynes, 2011). In a parallel cross-cultural study, mothers of firstborn 20-month-olds from seven countries (Argentina, Belgium, Israel, Italy, Japan, South Korea, and the United States) completed all the same procedures (Bornstein et al., 2007). The Big Five factor structure was extracted when the cross-cultural personality data were analyzed, and again the Big Five were found to relate differently to diverse parenting cognitions. Likewise, C. L. Smith et al. (2007) examined longitudinal relations among maternal personality, emotional expressions, and parenting sensitivity and intrusiveness: Conscientiousness and Agreeableness were positively associated with observed positive emotional expressions at T1, and Agreeableness, Openness to Experience, and Extraversion at T1 were positively related to positive emotional expressions at T2. Maternal positive emotional expressions at T1 and T2, in turn, were associated with more sensitive parenting observed at T3. In addition to first-order associations between maternal personality and maternal parenting, other factors (e.g., emotional expressiveness) may constitute possible pathways for explaining second-order relations between personality and parenting. Maternal personality in the normal range is often neglected in parenting research, but it appears to be operative in everyday parenting. Other specific features of normal personality favorable to good or poor parenting likely include empathic awareness, predictability, nonintrusiveness, and emotional availability. Lack of self-centeredness and adaptability might also be especially pertinent to parenting; so, parents who are more focused on relieving their own discomfort may resort to insensitive parenting strategies, such as minimization and punishment, and become dysregulated and try to end a child's distress abruptly (Fabes, Poulin, Eisenberg, & Madden-Derdich, 2002). Some states of mind, perhaps in the normal range but still troubling, are regularly identified with poor parenting (Whisman & Baucom, 2012; Zahn-Waxler, Duggal, & Gruber, 2002). Mood is one, and mood states predict parenting self-efficacy (Leerkes & Burney, 2007; Porter & Hsu, 2003; C. M. Weaver, Shaw, Dishion, & Wilson, 2008). Stress, in its many forms, has a rich history of untoward influence on multiple aspects of well-being in parenting (Crnic & Low, 2002; Deater-Deckard, 2004). The upbringing of children is highly emotional for both parents and children (Pomerantz, Wang, & Ng, 2005). Adult adaptability is likely vital, not only in the first few months, when infants appear unpredictable and undifferentiated, but through adolescence. Mothers and fathers report increased parenting stress across their child's transition to adolescence, and stress seems to arise from parent-adolescent interactions rather

than qualities of the parent or the child per se (Putnick et al., 2010). Stress, as in coping with daily hassles, predicts less maternal positivity (Crnic & Low, 2002); for example, depressed mothers experience more parenting stress (Coyl, Roggman, & Newland, 2002), and stress likely negatively affects parenting through disruptions in the parent-child relationship (Ciciolla, Crnic, & West, 2012; Crnic, Gaze, & Hoffman, 2005). Parenting stress is related to specific parenting behaviors (Putnick et al., 2008), and the stability of stress over time likely maintains adverse influences (Östberg, Hagekull, & Hagelin, 2007). Stress is also commonly related to diminished psychological well-being in parenting (Deater-Deckard, 1998). Abidin (1992) conceptualized parenting stress as resulting from many different relationships and contexts. Parents who are more stressed in the parenting role are more likely to behave in an authoritarian manner (Tan, Camras, Deng, Zhang, & Lu, 2012), whereas parents who report low levels of stress are more likely to be authoritative (Woolfson & Grant, 2006). Less stressed parenting is linked with better parent-adolescent communication (Joshi & Gutierrez, 2006), whereas more stressed parenting is associated with harsh discipline (Anjum & Malik, 2010; Fang, Wang, & Xing, 2012; Pinderhughes, Dodge, Bates, Pettit, & Zelli, 2000) and, under chronically high conditions, even with child abuse (Taylor, Guterman, Lee, & Rathouz, 2009). It has been argued that self-centered parents would be less likely to put children's needs before their own (C. P. Cowan & Cowan, 1992; Dix, 1991), so that women who are more preoccupied with themselves, as measured by physical and sexual concerns, likely display less effective parenting (Grossman, Eichler, & Winikoff, 1980). Still other possibly untoward characteristics include “parenting perfectionism”—the degree to which parents hold excessively high standards (M. A. Lee, Schoppe-Sullivan, & Kamp Dush, 2012; Snell, Overbey, & Brewer, 2005). Whereas psychological well-being is associated with parenting competence (Fujiwara, Okuyama, & Izumi, 2012), psychological distress (anxious or depressive symptomatology short of outright psychopathology) appears to undermine parenting. Maternal anxiety predicts self-reported diminished warmth toward children (Drake & Ginsburg, 2011), and anxious mothers may be disengaged or intrusive and overstimulating (Feldman, Granat, Pariente, Kanety, Kuint, & Gilboa-Schechtman, 2009; Kertz, Smith, Chapman, & Woodruff-Borden, 2008; for a meta-analysis, see van der Bruggen, Stams, & Bogels, 2008). For example, mothers who are anxious or angry in response to their infant's crying may misread these distress cues or believe that crying is a nuisance that interferes with their own goals, and so are unlikely to respond sensitively when their infants are distressed because they prioritize their own needs over those of their infants. Whether fleeting, as in response to transient economic circumstances or even the birth of the baby, or chronic, depression affects parenting adversely (Field, 2010; Lovejoy, Graczyk, O'Hare, & Neuman, 2000; Murray, Halligan, & Cooper, 2010; Zahn-Waxler et al., 2002). Epidemiologically, postnatal depression is surprisingly common with a prevalence of 10% to 15% (Gavin et al., 2005). Heron and colleagues (2004) even found that 44% of women with postpartum depression (and 36% of the women with postpartum anxiety) did not experience depression (or anxiety) in pregnancy. Depressed mothers fail to experience—and convey to their children—much happiness with life. Depression's associated mood disturbance, worry,

and rumination compromise mothers' ability to attend, diminish their responsiveness, and discoordinate their interactions with infants and children (Dix & Meunier, 2009; Manian & Bornstein, 2009; Murray et al., 2010; Pearson et al., 2012; Stein et al., 2012). Mothers who present with depression likely show increased negative affect and cognitions, apathy and lack of energy, and decreased engagement with children (Campbell, Matestic, von Stauffenberg, Mohan, & Kirchner, 2007; Kertz et al., 2008; Mertesacker, Bade, Haverkock, & Pauli-Pott, 2004; Niccols & Feldman, 2006; van Doesum, Hosman, Riksen-Walraven, & Hoefnagels, 2007). Overall, meta-analysis of this literature points to small but significant associations between maternal depression and negative (coercive, intrusive) and disengaged (withdrawn, uninvolved) parenting (Lovejoy, Graczyk, O'Hare, & Neuman, 2000). As with other psychological symptomatology, the effects of depression are not exclusive to mothers. Prenatal and postpartum depression manifest in about 10% of men (Paulson, Bazemore, & Sharnail, 2010); postpartum paternal depression is predicted by maternal depression (Goodman, 2004; Wee, Skouteris, Pier, Richardson, & Milgrom, 2011); and depression exerts untoward effects on paternal parenting as it does on maternal parenting (R. N. Davis, Davis, Freed, & Clark, 2011; Wilson & Durbin, 2010). Happily, most cases of postpartum depression remit; unhappily, treatment of depression in those who do not is less than successful in repairing parent-child relationships (Forman et al., 2007). Outright psychopathology, such as mental illness, phobias, substance abuse, and antisociality, seriously impairs thinking, affect, and behavior, and consequently parenting cognitions and practices. Vesga-López et al. (2008) estimated prevalences of postpartum psychopathologies to range from 12% for substance use to 15% for mood disorders. Fewer studies of psychopathology and parenting populate the literature (Berg-Nielsen, Vikan, & Dahl, 2002). However, mothers diagnosed with major psychopathologies display detached parenting, spend less time with their children, show less affection, and provide less structure (Brook, Brook, Ning, Whiteman, & Finch, 2006; Champion et al., 2009; Gerdes et al., 2007); and mothers who engage in antisocial behavior are, on the one hand, less likely to employ optimal parenting practices (Jaffee, Belsky, Harrington, Caspi, & Moffitt, 2006) and, on the other hand, more likely to demonstrate hostility toward children (Bosquet & Egeland, 2000) or even abuse them physically (Kim-Cohen et al., 2006a). Intergenerational Transmission and Family of Origin Winnicott (1948/1975) and Spitz (1965/1970) detected the roots of aggressive, impulsive, immature, self-centered, and self-critical caregiving in parents' own experiences growing up. Likewise, Bowlby (1969) and successive attachment theorists subsequently proposed that, arising out of their early interpersonal experiences with caregivers, young children develop “internal working models” of their caregivers that incorporate both sides of the caregiver-child relationship. Attachment theory asserts that, on the basis of repeated experiences of characteristic patterns of interaction, children develop expectations regarding the nature of social interactions that endure long after they depart their family of origin. Representations of interactions with parents thus model individual differences in attachment security in childhood and reverberate at later ages in associations with others (Fraley & Shaver, 2000; Mikulincer,

Shaver, & Pereg, 2003) as, for example, in the quality romantic relationships (Dekovic & Meeus, 1997) and in own parenting. For instance, evidence from nonhuman animal research shows that the experience of being neglected or abused can affect infant brain development, with the result that inadequate parental behavior is exhibited in adulthood (Champagne, 2008; Maestripieri, 2005; Numan & Insel, 2003; Numan & Stolzenberg, 2009). Mothers with lower maternal care in childhood have larger gray matter volumes in key brain areas related to parenting and show lower activation there in response to infant cries; members of high-risk families fail to show patterns of activation in brain regions associated with empathy and collaborative behavior (Atkinson, 2012; P. Kim et al., 2010b). Intergenerational continuities in parenting have been detected in many different mammalian species, including rats, rhesus monkeys, and human beings (Conger, Belsky, & Capaldi, 2009; Jensen & Champagne, 2012). A lay belief supported by an emerging science is that the way parents rear their offspring results to a significant degree from the child-rearing parents experienced when they were growing up (Conger, Schofield, & Neppl, 2012). In animals, the quality of parental care received in infancy is a significant predictor of offspring and even grand-offspring parental behavior, suggesting that intergenerational effects persist (Jensen & Champagne, 2012). Parenting styles are often similar from one generation to the next. In laboratory rats, mothers, daughters, and granddaughters lick and groom their litters at similar frequencies (Champagne & Meaney, 2007), and cross-fostering studies support the hypothesis that the behavioral transmission of maternal behavior is not dependent on the transmission of specific DNA variants (Champagne et al., 2003). A mother's reported family-of-origin emotional expressiveness is important to understanding the manner in which she socializes emotion in her child. Baker and Crnic (2005) examined the role of mothers' reports of their family histories of emotional expressiveness for the manner in which they socialized emotion in their children: Systematic relations between reports of family-of-origin expressiveness and current parenting emerged for certain forms of expressiveness. Through intergenerational transmission, via interlocked genetic and experiential pathways, purposefully or unintentionally, then, one generation (G1) appears to influence the parenting beliefs and behaviors of the second generation (G2) and thus the experiences and even child rearing of the third generation (G3). Fraiberg, Adelson, and Shapiro (1975) once referred to these inspirations as “ghosts in the nursery.” In two early studies, Ruoppila (1991) reported significant correlations between grandparental and parental child rearing in a Finnish sample, and Vermulst, de Brock, and van Zutphen (1991) documented similarities in parental functioning across generations in a Dutch sample. Maritally dissatisfied couples are more likely to have had unhappily married parents (Amato & Booth, 2001); marital violence in the family of origin tends to repeat in the successive generation (Stith et al., 2000); and when parents abuse their children, their children are at risk of repeating abuse as parents with their own children (Cicchetti, Toth, & Maughan, 2000; Newcomb & Locke, 2001; Pears & Capaldi, 2001). Kovan, Chung, and Sroufe (2009) provided a “longitudinal” illustration of intergenerational transmission when they recorded interactions of parents (G1) and their 2year-olds (G2) and then waited and recorded interactions of those 2-year-olds as parents (G2) of their own 2-year-olds (G3). Even accounting for confounds, a relatively strong

correspondence emerged in parenting between generations. A variety of studies now documents similarities across generations for harsh parenting (Capaldi, Pears, Patterson, & Owen, 2003; DiLillo & Damashek, 2003) and poor supervision (C. A. Smith & Farrington, 2004) as well as for positive (Hofferth, Pleck, & Vesely, 2012; Thornberry, 2005; Thornberry, Freeman-Gallant, Lizotte & Krohn, 2003) and constructive caregiving (Z. Chen & Kaplan, 2001; Kerr, Capaldi, Pears, & Owen, 2009). Using data from grandparents (G1) and parents (G2), Bailey, Hill, Oesterle, and Hawkins (2009) examined continuity in parental monitoring and harsh discipline across generations. Parental monitoring and harsh discipline demonstrated continuity from G1 to G2. Thus, parenting in G1 vis-à-vis G2 predicts parenting in G2 toward G3. Attachment is lifelong (Mikulincer & Shaver, 2010), and the Adult Attachment Interview (AAI; George, Kaplan, & Main, 1985) assesses an adult's (G2) model of his or her own relationships with his or her parents (G1). Unsurprisingly, adult attachments are conceptualized as relatively stable patterns of feelings, thoughts, and behaviors that characterize a person's current close relationships (Shaver & Mikulincer, 2002). For example, attachment avoidance reflects discomfort with closeness and intimacy. People who avoid try to minimize their distress, are uncomfortable depending on others or having others depend on them, and are reluctant to disclose feelings or information to relationship partners that might suggest vulnerability (Mikulincer & Shaver, 2010; Shaver & Mikulincer, 2002). Adult attachment insecurity, particularly avoidance, is associated with more negative responses to infant distress in females (Leerkes & Siepak, 2006) and less sensitive and comforting responses to children in mothers (Edelstein et al., 2004). Strong predictive links have emerged between the G2 mother's AAI classification vis-à-vis parents in G1 and the G3 child's attachment with the G2 mother (P. K. Smith & Drew, 2002). G2 mothers who report having secure and realistic perceptions of attachments to their G1 mothers, for example, are themselves more likely to behave sensitively with their G3 children and have securely attached G3 children (Cummings & Cummings, 2002). Van IJzendoorn (1995) reviewed AAI studies and reported 75% concordance between the parent's autonomous/nonautonomous classification and a child's secure/insecure classification in Ainsworth's Strange Situation. Indeed, some claim that G2 maternal attachment to her primary G1 caregiver is a better predictor of her parenting skills than G2 representations of her G3 child or experiences directly with the G3 child (Biringen, Matheny, Bretherton, Renouf, & Sherman, 2000). Thus, adults' caring for their offspring may have roots in their own early-life experiences with primary caregivers such that people who experienced quality parenting as children may express stronger compassion and altruistic behaviors towards others. This conclusion is supported by the continuity of parent-infant attachment styles into adult relationships (Fraley, 2002). Parents appear to draw on the legacy of their families of origin for models of parenting, often emulating the same patterns of parenting and couple interaction they were exposed to as children, so mothers' positive memories of their family of origin's coparenting relationship are associated with their own supportive approach (Stright & Bales, 2003). For obvious reasons of longitudinality, the field of parenting in human beings has made only modest progress on

what is an extraordinarily challenging problem: namely, establishing that patterns of parenting in G1 are internalized and carried forward to be reactivated in G2. Just collecting two or three generations of data is formidable, so it is to be expected at this stage that the field is far from settled on the mediational processes by which parenting transmits intergenerationally. Both genetic and experiential pathways of various types have been proffered. For example, a parent's experiences with and memories of his or her own parents may reverberate in his or her own parenting via direct (possibly ascribable to G2's observational learning of G1 parenting) or indirect (experiencing G1 parenting as a child affects mature behaviors in interaction with others, including one's own children) means (Campbell & Gillmore, 2007; Capaldi et al., 2003; van IJzendoorn, 1992). Whether the epigenetic activation of maternally related genes is transmitted across multiple generations of female offspring is not known, but cycles of parenting clearly repeat across multiple generations (Bower-Russa, Knutson, & Winebarger, 2001; Champagne, 2010; Gonzalez, Lovic, Ward, Wainwright, & Fleming, 2000; Kaffman & Meaney, 2007; Veenema, 2009). A prospective, intergenerational study considered multiple influences on fathers' constructive parenting. Fathers in G2 were recruited as boys on the basis of neighborhood risk for delinquency, and they were assessed through early adulthood. The fathers' parents (G1) and the G2 mothers of G3 children were also studied. A multiagent, multimethod approach measured G1 and G2 constructive parenting (monitoring, discipline, warmth, and involvement) and G2 positive adolescent adjustment. Path modeling supported direct transmission of G1 constructive parenting of G2 in late childhood to G2 constructive parenting of G3 in middle childhood. G1 parenting also indirectly influenced G2 parenting through G2 positive adjustment. G1 parenting influenced G2 parenting in both early and middle childhood of G3. Positive father involvement in terms of caring for the child and sharing decision making when sons are young is associated with warmer and more positive parenting of their young children by the same sons as parents themselves in young adulthood; even in a high-risk sample, young men whose fathers were involved became better fathers to their own children (Hofferth et al., 2012). These intergenerational continuities in parenting persisted, even when additional influences were considered. Transmission pathways of constructive parenting are maintained, in part, by engendering positive adjustment in offspring (Kerr et al., 2009). Two important questions in this literature ask, first, whether similar intergenerational continuities obtain for positive as well as negative aspects of parenting and, second, what types of processes or mechanisms might account for these associations (Conger et al., 2012). Using observer reports of parent-child interactions in the home, Neppl, Conger, Scaramella, and Ontai (2009) evaluated G2 parenting of their G3 children. With regard to the first question, these researchers found continuities between G1 and G2 harsh parenting (angry, hostile, and uncaring behaviors) and between G1 and G2 positive parenting (interest, concern, and clear communication). However, G1 positive parenting did not predict G2 harsh parenting after controlling for harsh parenting in G1. These findings help to identify specific mediating pathways that might explain intergenerational continuities in both styles of parenting, and they are consistent with earlier research demonstrating that negative parenting behaviors in G1 are related to the same behaviors in G2 through more general

conduct problems of G2 parents (Caspi & Elder, 1988). Neppl and her colleagues (2009) also showed that G1 harsh parenting predicted G2 externalizing or antisocial behaviors, which then led to G2 harsh parenting. That is, when G2 youth experienced harsh parenting as adolescents, they were more likely to engage in antisocial behaviors during late adolescence and early adulthood, and these antisocial tendencies were directly related to their harsh parenting behaviors with their children. The authors proposed that positive or constructive parenting by G1 would be related to the same type of parenting in G2 through the competencies that positive parenting promotes during childhood and adolescence. G2 academic attainment mediated the relations between G1 and G2 positive parenting. Externalizing behaviors did not mediate continuity in positive parenting, and academic attainment did not mediate continuity in harsh parenting. It appears that these intergenerational continuities are explained by processes or mechanisms specific to the type of parenting being considered. Conger (2013) showed that continuity in harsh parenting is partly a function of the partners that one chooses. People who have been harshly parented in childhood are more likely to partner with those who parent harshly, which in turn increases the parent's own risk of harsh parenting. In brief, specific styles of parenting demonstrate continuity across generations. A special section of Developmental Psychology (Conger et al., 2009) was devoted to the issue of the intergenerational transmission of parenting, including evidence regarding the intergenerational transmission of positive parenting and developmental mediators that seem to be involved in that transmission. Evidence for intergenerational continuity in parenting has emerged as robust across diverse study samples, types of measurement, lengths of time, and after consideration of a variety of confounders. These studies also raise important questions about genetic and epigenetic processes that might contribute to similarities and dissimilarities in parenting across generations. Epigenetics and Parent Experiences Epigenetics constitute regulatory layers “on top of” DNA that may be experiential and modulate gene expression. For example, male mice learned to fear the smell of cherry blossoms (through associative conditioning with foot shocks); they were then mated with females. Their offspring, raised to maturity without ever having been exposed to the scent of cherry blossoms, demonstrated fearfulness the first time they were exposed to cherry smell and had a lower threshold for cherry blossom detection, and their brains had more neurons devoted to the scent. Moreover, the pups of females artificially inseminated with sperm of fearconditioned father mice showed similar effects (Dias & Ressler, 2014). Thus, characteristics outside DNA appear to be inherited from parents (see also Vassoler, White, Schmidt, SadriVakili, & Pierce, 2012).

Summary Likely most proximal to parenting are determinants in the parent, including biological characteristics of the parent (such as genetic endowment, hormonal activity, and nervous system structure and function) as well as psychological characteristics (such as the parent's gender, age and stage of life, health status, cognitions, personality, family of origin, and

perhaps parents' own idiosyncratic experiences). All of these factors (and others in the child and context detailed in the next sections) need to be taken into consideration when attempting to divine sources of individual differences in parenting cognitions and practices in mothers, fathers, and other child caregivers.

Determinants in the Child When rat dams are given a choice between a chamber that has been associated with pups and one where they may receive an injection of cocaine, there is a period of time after parturition when new dams prefer the company of pups and succumb to power of offspring, spurning even the allure of narcotics (Mattson, Williams, Rosenblatt, & Morrell, 2001; Pereira, Seip, & Morrell, 2008; Seip & Morrell, 2007). Thinking about parent-child relationships naturally highlights parents as agents of child socialization, and there is ample evidence, especially in the early years, that, between parents and children, children have little agency and parents exert more sway (Kochanska & Aksan, 2004; Maccoby, 1992; Vygotsky, 1978). There is initially asymmetry between parent and child contributions to parenting, but children from infancy affect parents. Parenting is a two-way street, and it has long been appreciated that development in children is a catalyst for transformations in parenting (Baldwin, 1946) and that children actively select, modify, interpret, and create their own environments, including their parenting (Bell, 1968; Scarr & Kidd, 1983). As the child ages, the parent-child relationship becomes more mutual and less hierarchical, providing adolescents and adult children with opportunities that help to redefine their relationships with their parents (Bornstein, Jager, et al., 2012). Contemporary theorizing readily acknowledges the effects of children on parents (Bell, 1968, 1970; Sameroff, 2009). Often what parents think or do depends on many child factors, and subtle as well as not-sosubtle biological and physical, psychological, and social characteristics of children influence parenting (Hodapp & Ly, 2005; Karraker & Coleman, 2005). Child effects answer the manifest behavioral adjustments parents make to children's age and gender and appearance, temperament, and activity. Child effects have several construals; from attachment theory, they signal need for care or alternatively they prompt or cause parenting (Sroufe, 1985). Some child effects are universal and common to all children; others are unique to a particular child or child-related situation. In early childhood, maternal sensitivity is in part defined by the reciprocal nature through which parents respond to their children, suggesting flexibility in parenting practices to adjust to minute-by-minute demands of a situation (Ciciolla et al., 2012). The extent to which parents are sensitive depends on how well they are able to mold their responses and interactions to the needs of the child (Ainsworth, 1969; Leerkes, Blankson, & O'Brien, 2009); thus, determinants in the parent and in the child interact. From the start, the birth of a child stirs the emotions and rivets the attention of adults. By their very coming into existence, children alter the sleeping, eating, and working habits of their parents; and they change who parents are and how parents define themselves. Moreover, parent and child activities are characterized by intricate patterns of mutual attunement and transaction (Bornstein, 2009, 2013a). Infants cry to be fed and changed, and when they wake, they let

parents know they are ready to play and to learn; through social referencing and attachment relationships, infants and young children regularly deploy parents as agents. Many parenting initiatives are proactive; very often, however, parents behave reactively. When Anderson, Lytton, and Romney (1986) experimentally paired conduct-disordered boys with mothers of normal boys, the mothers behaved negatively. Adolescent problem behaviors may more powerfully determine parenting practices than parenting practices provoke subsequent adolescent problem behaviors (Huh, Tristan, Wade, & Stice, 2006). Bell (1968; Bell & Harper, 1977) was among the first developmental theorists to emphasize the key role that bidirectional effects play in child socialization. By virtue of their unique characteristics and propensities—genetic endowment, physical features, state of arousal, perceptual awareness, cognitive capacity, emotional expressiveness, and individuality of temperament and personality—children actively contribute to the parenting they receive. Children influence which experiences they will be exposed to, and they interpret and appraise those experiences and so (in some degree) determine how their experiences affect them. Even within the same family and home setting, parents do not behave toward their different children in the same way, parenting is not perceived by different children in the same way, and parenting does not affect different children in the same way (Suitor et al., 2009; Turkheimer & Waldron, 2000). Some researchers have pointed to the relatively important role of children's perceptions of parenting in examining effects of parenting (Barry, Frick, & Grafeman, 2008). For example, Demo, Small, and Savin-Williams (1987) found that children are influenced more by their perceptions of parental practices and less by the actual practices. The mere presence of offspring can be efficacious. When Kenkel et al. (2012) exposed reproductively naive male prairie voles (Microtus ochrogaster) to infants or control manipulations and measured plasma concentrations of OT, they found that simple pup exposure increased activation of neurons that stained for OT. Becoming or being a parent affects adults, females and males alike. Males in the biparental marmoset species Callithrix jacchus engage in high levels of parenting and express enhanced circulating reproductive hormones, such as arginine vasopressin. The brains of first-time and experienced marmoset fathers have greater abundance of arginine vasopressin V1a receptors and greater density of V1a receptor-labeled dendritic spines on pyramidal neurons in prefrontal cortex than nonfathers (Kozorovitskiy, Hughes, Lee, & Gould, 2006). Woller et al. (2012) examined the release of several reproductive neurocrines, including OT and PRL, in cultured explants of the hypothalamus of paternally experienced male marmosets compared to naive paternally inexperienced males. OT and PRL levels were higher than levels found in inexperienced males, suggesting that paternal experience prompts secretion of neurocrines in a male biparental primate.

Biological Characteristics Physical features of children likely affect parents everywhere, perhaps in similar ways. By 18 to 20 weeks of gestation in their first pregnancy (16–18 weeks in repeated pregnancies), fetuses are felt to move in utero (“quickening”), a significant marker in the lives and psyches of the child's parents (Cunningham et al., 2010). Newborns have large heads dominated by disproportionately large foreheads, widely spaced relatively large eyes, small snub noses,

exaggeratedly round faces, and weak chins. The ethologist Lorenz (1935/1970) hypothesized that these physiognomic Kindchenschema incite adults to express nurturant reactions. Adults are biased to respond quickly to human infant faces in positive ways (Esposito et al., 2015; Parsons, Young, Kumari, Stein, & Kringelbach, 2011; Senese et al., 2013) and are strongly motivated to caregive by them (Glocker et al., 2009a), just as human infant faces increase activation in select brain regions associated with empathy and responsiveness (Caria et al., 2012; Glocker et al., 2009b). Cuteness in infant faces modulates mother-infant interaction (Langlois, Ritter, Casey, & Sawin, 1995), and cuter babies are rated as more friendly, cheerful, likable, healthy, and competent (Casey & Ritter, 1996; Karraker & Stern, 1990; Ritter, Casey, & Langlois, 1991; Stephan & Langlois, 1984). A female's willingness to adopt a baby (Volk & Quinsey, 2002) and a mother's sensitivity toward a baby (Langlois et al., 1995) in part reflect facial cuteness. Viewing a picture of one's own child activates brain areas associated with reward and motivation (Mascaro, Hackett, & Rilling, 2013). Idiosyncratic characteristics of individual children are no less stimulating to parents. Every child is an original, and general developmental functions in children unfold in the context of wild individual variation. Some individual differences are rooted at deep biological levels. For example, a polymorphism of the DRD4 gene is associated with disorganized attachment (Gervai et al., 2005; Lakatos et al., 2000, 2002). Among carriers of the 7-repeat DRD4 allele, there is no relation between quality of maternal communication and infant attachment; however, a strong relation exists between maternal disrupted communication and attachment in infant carriers of the DRD4 7-repeat genotype. That is, an individual-differences genetic trait in the child moderates the relation between maternal caregiving and infant attachment (Gervai et al., 2007). Similarly, low maternal sensitivity in infancy predicts higher levels of mother-reported externalizing behavior problems at 2 to 3 years of age but only if infants carry the 7-repeat allele of the DRD4 gene (Bakermans-Kranenburg & van IJzendoorn, 2006). Caspi and colleagues (2002, 2003) reported that the relation between childhood maltreatment and later psychological maladjustment is moderated by genetic factors. The functional polymorphism of the regulatory region of the monoamide oxidase A (MAOA) gene moderates the relation between early maltreatment and later antisocial behavior, and a regulatory polymorphism of the serotonin transporter (5-HTT) gene moderates the effect of early maltreatment on adult depression (see also Foley et al., 2004; Kaufman et al., 2004; Kim-Cohen et al., 2006b). O'Connor et al. (1998) identified two groups of adoptees: one at genetic risk for antisocial behavior (their biological mothers had a history of antisocial behavior) and the other not at risk. At several points during the adoptees' childhood, O'Connor et al. assessed the children's characteristics and the adoptive parents' child-rearing methods. Children carrying a genetic risk for antisocial behavior were more likely to receive negative socialization from their adoptive parents. Jaffee, Caspi, Moffitt, and Taylor (2004) found that 5-year-old twins' experience of corporal punishment was mediated by the same genetic factors that predisposed them to aggressive and oppositional behavior; children's antisocial behavior evoked physical maltreatment. Likewise, Pener-Tessler et al. (2013) learned that boys' self-control mediated the association between their serotonin transporter gene polymorphism (5-HTTLPR) genotype and their mothers' positivity (warmth, autonomy support, responsiveness) toward them: Boys who carried the short allele of 5-HTTLPR exhibited higher levels of self-control and received

higher levels of positive parenting. Other biological characteristics of children, even if seemingly hidden, appear to affect parents. Vagal tone indexes heart rate and breathing changes and, as a physiological marker of emotion regulation, estimates parasympathetic nervous system influences on the heart. Mothers are concordant with their children in vagal tone (Bornstein & Suess, 2000), and higher vagal tone (more parasympathetic regulation) in children predicts more supportive and less restrictive mothering, even controlling for earlier mothering (Kennedy, Rubin, Hastings, & Maisel, 2004). Normal development may be nonlinear in nature, stalling sometimes or even regressing temporarily (Bever, 1982; Harris, 1983; Strauss & Stavey, 1982). Parenting a child is thus akin to trying to judge a moving target, the ever-changing child developing in fits and starts at his or her own pace. Interest in the origins and expression of their child's individuality occupies a central position in parental thinking. Twin, adoption, and sibling studies indicate that child temperament and developmental competence are both instrumental in affecting the nature of parenting (Plomin, Reiss, Hetherington, & Howe, 1994). A persisting parenting challenge is that, at base, parents are constantly trying to divine what is inside their children's heads, what they want, what they know, how they feel, and what children will do next vis-à-vis the people and things around them. Thus, parents seem constantly in search of patterns, often inferring them on the basis of single transient events. For example, the ages at which individual children achieve a given developmental milestone typically vary enormously (some children say their first word at 9 months, others at 29 months), just as children of a given age vary dramatically among themselves on nearly every index of development (at 20 months, some toddlers speak 1 word, some 487; Bornstein et al., 2004). Of course, when and how their children talk, reach puberty, and so forth exercise strong psychological and behavioral impacts on parents.

Psychological Characteristics Individual-difference characteristics of children stimulate parenting beliefs and behaviors, which then influence children's adjustment and development. Goldberg (1977) usefully taxonomized three salient child characteristics that affect parents: responsiveness, readability, and predictability. Responsiveness refers to the extent and quality of child reactivity to stimulation; readability refers to the definitiveness of child behavioral signals; and predictability refers to the degree to which child behaviors can be anticipated reliably. Each child varies along each of these dimensions and manifests a unique profile of these characteristics (and others) that will influence parenting. For example, an “easily read” adolescent produces unambiguous cues that allow parents to recognize the adolescent's state quickly, interpret signals promptly, and thus respond appropriately. Age and Stage Some aspects of parenting are frequent or significant from the get-go and wane as children develop; others wax in importance over the course of child development. There is a general reduction in time devoted to parenting activities, especially caregiving, as children grow and develop and puberty distances adolescents from their parents. Parents of younger children (0–6

years) spend twice as much time in child care activities as do parents of older children (6–17 years; U.S. Bureau of Labor Statistics, 2013), and the sheer frequency with which children are exposed to parent cognitions and practices changes markedly from infancy to adolescence (Bradley, Corwyn, McAdoo, & García Coll, 2001). Child age predicts maternal reports of their own positive and negative parenting in a Canadian cohort study of children aged 4 to 11 years (Jenkins, Rasbash, & O'Connor, 2003) as it does parental reports as well as observations of parenting (Browne et al., 2012). As their susceptibility to peer influences peaks, adolescents experience a concomitant decline in susceptibility to parental influence, more and more determine how much time they spend with their parents, and the extent to which their parents are involved in and know about their lives outside the family (Bornstein, Jager, et al., 2012; Kerr, Stattin, Biesecker, & Ferrer-Wreder, 2003). Mothers and fathers alike report increased parenting stress across their child's transition to adolescence (Putnick et al., 2010), and adolescents and parents contest control over different types of issues as adolescents age (Smetana & Villalobos, 2009). Younger adolescents tend to endorse parental responsibility and care, whereas older adolescents tend to justify expression of rights with appeals to personal choice and rights (Cherney, 2010). Thus, adolescents come to consider some behaviors and actions to fall outside the bounds of parental regulation (Smetana, 2000), such as personal issues (e.g., hairstyle or favorite food), multifaceted issues (containing aspects of both personal and conventional domains), and prudential issues (involving potential harm to an actor). Parents, by contrast, continue to view these issues as within the scope of their authority, defining them as social conventions or matters of safety. These discrepant interpretations can lead to increases in everyday conflicts that ultimately serve as opportunities for adolescents and parents to renegotiate boundaries of authority (Smetana & Villalobos, 2009). Adolescents reportedly experience negative moods more often than either children or adults (Larson & Asmussen, 1991; Larson, Csikszentmihalyi, & Graef, 1980; Larson & Lampman-Petraitis, 1989), and the increase in negative affect during adolescence challenges parents of teenagers. One cross-sequential study that spanned ages 6 through 18 reported that positive aspects of parenting decreased markedly through middle adolescence even though parent-child communication changed little (Loeber et al., 2000). As a result, parents need to calibrate their reactions to adolescents' emotions. Young children's interactions with their context are normally controlled by their parents, whereas older children and adolescents may engage more closely with other people, institutions, and physical aspects of their contexts (Leventhal, Dupéré, & Shuey, 2015). For example, children's developmental status makes different neighborhood social processes more or less salient for their parents (Skinner et al., 2014). Neighborhood social cohesion may be more important for parents when their children are young because of parenting demands, but parents may become more focused on neighborhood disorder as their children age and have greater direct exposure to neighborhoods. Parents adjust relative to both child age and stage and child capacity and performance (Bellinger, 1980). Children's achieving certain milestones—standing upright and walking, for example—completely alters the nature and quality of adult caregiving (Campos et al., 2000), just as do many other rites of passage, such as children going to school for the first time,

receiving communion or getting bar or bat mitzvahed, driving and dating, attending college, getting married, and all the events in between. With each such child development, parenting changes in some corresponding ways. Parents talk to the walking as opposed to the crawling toddler differently, just as they interact with the pubertal adolescent differently from the prepubertal. It is an understatement to say that adolescents' initiation of dating and sexual relationships have consequences for parents. Sensitive parents tailor their parenting to match their children's developmental progress (Adamson & Bakeman, 1984; Carew, 1980), for example, by providing more didactic experiences as children become more competent (Bornstein & Tamis-LeMonda, 1990; Bornstein et al., 1992; Klein, 1988). Mothers of infants in many cultures use affect-laden speech, but as their young children achieve more sophisticated levels of motor exploration and cognitive comprehension, mothers increasingly orient, comment, and prepare them for the world outside the dyad by infusing maternal speech to children with increasing amounts of information (Bornstein et al., 1992). The mean length of mothers' utterances tends to match the mean length of those of their 12- to 32-month-olds (McLaughlin, White, McDevitt, & Raskin, 1983). Childhood is change—development involves rapid growth in biological, mental, emotional, and social spheres; understanding, anticipating, and responding to these dynamic changes continuously challenge parents. Parents need to know about and keep vigilant to all the complications and subtleties associated with their child's development. With each child advance, parenting changes in some corresponding ways. Individual parents seem to maintain their standing relative to one another from day to day: Parents who talk to their children more on Monday likely talk to their children more on other days of the week. Over longer periods, of course, group mean levels in parenting change, and they certainly do so in response to children's development. Thus, stability does not negate age-appropriate parenting. Nonetheless, keeping an eye on a toddler in the park when the child is beginning to walk and giving a cell phone to a teen who is beginning to drive a car appear superficially different but may be conceptually similar in the sense of reflecting continuity in parental monitoring. As Holden and Miller (1999, p. 243) concluded, “the nature of child rearing is simultaneously enduring and different.” Birth Order Survival rates for primate infants are higher among experienced multiparas compared with primiparas (Hrdy, 2009). For example, 63% of baboon infants of experienced mothers survive versus 29% of infants born to first-time mothers (Altmann, Hausfater, & Altmann, 1988). In human beings, multiparas report higher self-efficacy than primiparas (Fish & Stifter, 1993). Although a substantial proportion of developmental science has been constructed on motherfirstborn relationships (Hoffman, 1991), roughly 80% of families in the United States will have more than one child (Dye, 2010), and most children in the United States grow up with siblings (Kreider & Ellis, 2011). Family structure, especially parity, has been thought to play a signal role in the development and expression of individual differences in parenting at least since Francis Galton (1874) and Alfred Adler (1928). A dramatic change in family dynamics takes place when a second baby is born into the family, and parenting differs in some ways for first-

versus laterborns (J. Dunn & Plomin, 1990; Sulloway, 1996). Mothers engage, respond, stimulate, talk, and express positive affection more to first- than to laterborns, even when firstand laterborns show no differences in their behavior, indicating that some parenting does not (necessarily) reflect child behavior but parity (Belsky, Gilstrap, & Rovine, 1984). Combined with variation in genetic makeup, within-family variation in parental thinking and treatment is a potent factor in accounting for why children in the same family differ from one another (Caspi et al., 2004). This line of thinking culminates in parental differential treatment (Solmeyer, Killoren, McHale, & Updegraff, 2011) of siblings, which occurs when siblings are recipients of different child-rearing cognitions and practices. Parental differential treatment follows from children's individual characteristics (e.g., gender, age, personality, or special needs) and because of parental characteristics and family functioning (e.g., favored child versus scapegoat; McHale & Pawletko, 1992; Mekos, Hetherington, & Reiss, 1996). Thus, a parent may feel more affinity with or be more effective with one child than another. Parental differential treatment implies that parents do not have the same attitudes or act towards all of their children simularly. Indeed, there is research that supports the supposition that parents' cognitions and practices are less favorable in families with greater numbers of children (Furman & Lanthier, 2002). Gender Although there is evidence that parenting girls and boys is surprisingly similar in many ways (Bornstein et al., 2015; Hyde 2005, 2014; Lytton & Romney, 1991), child gender broadly organizes many parent descriptions, impressions, and expectations of children from the start of life (Bornstein, 2013a): Parents tend to interact with children differently by child gender. Classic “Baby X” studies (where the gender of the infant is not known to study participants) in the United States have shown that unidentified infants are judged more frequently to be male (Condry & Condry, 1976; J. Z. Rubin, Provenzano, & Luria, 1974; Seavey, Katz, & Zalk, 1975) and that parents conceive of and behave toward infants differently depending on whether they think they are interacting with a girl or a boy (Seavey et al., 1975; Sidorowicz & Lunney, 1980). Newborn nurseries provide uniformly gender-appropriate blankets, accessories, and the like; baby showers carefully respect child gender; and infants are fastidiously dressed in gender-appropriate clothing (Shakin, Shakin, & Sternglanz, 1985). Parents have been reported to purchase gender-stereotyped toys for their children within a few months of the child's birth —prior to when children could express gender-typed toy preferences themselves (Pomerleau, Bolduc, Malcuit, & Cossette, 1990). Mondschein, Adolph, and Tamis-LeMonda (2000) found that mothers of 11-month-old boys would overestimate how well their babies could crawl down a sloped pathway, whereas mothers of 11-month-old girls would underestimate how well their babies could do, even when subsequent tests of crawling ability on the sloped path revealed no sex differences in infant crawling. Later, parents speak more frequently about emotions with their girls than with their boys, they speak about emotions differently, and they differentially reward and punish different emotions in daughters versus sons (Garside & Klimes-Dougan, 2002). Mothers respond differently to their “difficult” sons than to their “difficult” daughters, even when boys and girls do not differ on an independent measure of difficulty (Else-Quest, Hyde, Goldsmith, & Van Hulle, 2006). Daughters are apt to engage in

more relational and less autonomous play than boys, possibly fostering more sensitive responding from mothers (Ruble, Martin, & Berenbaum, 2006). Continuing on in development, although parents encourage participation in sports with no gender differences, parents are more likely to encourage participation in youth organizations, religious activities, and summer situations with sons than with daughters and are more likely to encourage participation in art programs with daughters than with sons (MacDonald & Parke, 1984). Parents also tend to give adolescent boys more decision making leeway (Daddis & Smetana, 2005), whereas adolescent girls report more stringent rules regarding socializing with friends, and caregivers are more likely to become involved in peer relationships because of misbehavior of their sons than of their daughters (Greene & Way, 2005; Way & Greene, 2006). Parents also appear to be more consistent in the ways they socialize their daughters than in the ways they socialize their sons (McGue, Elkins, Walden, & Iacono, 2005). For example, Kagan and Moss (1962) reported a higher cross-time correlation (r = .77) for maternal control of girls than boys (r = .24). Gender stereotypes influence parents' reports of adolescent problems; consequently, parental attributions reflect their differential expectations for boys versus girls (Bornstein, Putnick, et al., 2011; Bugental & Happaney, 2002; Lansford et al., 2011). Although conflicts with parents reportedly increase in intensity as warmth decreases across adolescence, these changes appear to be greater for girls than boys (McGue et al., 2005). Girls may spend more time with their mothers than do boys, becoming closer and more positively involved with them (Bornstein et al., 2008; Clarke-Stewart, 1973). Girls are recipients of greater warmth than boys (Atzaba-Poria & Pike, 2008), more positive parenting practices (Lloyd & Devine, 2006), and less harsh parenting and physical punishment (Lytton & Romney, 1991). In addition, girls are yelled at and smacked by their parents less often than boys are. Of particular relevance here is that a meta-analysis by Lytton and Romney (1991) found stronger effect sizes for parental physical punishment (strongly associated with parent rejection) for boys than for girls. Studies also have shown differences in the quality, pattern, and impact of mother-child and father-child interactions. All that said, the specific mechanisms that drive the influences of child gender on parenting cognitions and practices remain largely unknown (Ciciolla et al., 2012) as do the parenting dynamics of gender-same versus genderdifferent siblings (Jenkins et al., 2003). Health Status Child medical risk factors, such as low birthweight, prematurity, illness, developmental disability, and ASD, to name only a few, are associated with parenting stress (Eisengart, Singer, Fulton, & Baley, 2003; Goldberg & DiVitto, 2002; Hodapp, 2002; Kuhn & Carter, 2006; Saigal & Doyle, 2008; Zuckerman & Keder, 2015), and (unsurprisingly) the more severe the child's symptoms, the greater the parenting stress (Williford, Calkins, & Keane, 2007). Even multiple births pose problems (Damato, 2005; Lytton & Gallagher, 2002). Early developmental risk, such as prematurity and developmental delay, is associated with increased negative emotionality and behavior problems, less competent social interactions, and unclear emotional expression (Boström, Broburg, & Hwang, 2010; Feldman, 2007). In turn, these circumstances and behaviors are associated with greater parenting stress and more

psychological distress in the parents (Baker, Blacher, Crnic, & Edelbrock, 2002; Baker, Blacher, & Olsson, 2005), perhaps resulting in the lower levels of maternal sensitivity often reported in families of children with developmental challenges (Feldman, 2007; Fenning, Baker, Baker, & Crnic, 2007; Niccols & Feldman, 2006). Preterm children have difficulty regulating interpersonal engagements, as evidenced in their increased gaze aversion, decreased play, and lower levels of joint attention: in compensation, mothers of preterms are more active and directive (Goldberg & DiVitto, 2002; Winstanley et al., 2015). Similarly, children's problem behaviors affect parents: For example, children's cry-fuss/sleep problems contribute to maternal depression and stress, and repeated problems have the greatest impact (St. JamesRoberts, 2008). Mothers of children with developmental delay are less sensitive than mothers of typically developing children, especially under conditions of challenge (Ciciolla et al., 2012; Niccols & Feldman, 2006). If, as the Centers for Disease Control and Prevention's Autism and Developmental Disabilities Monitoring Network Surveillance Year 2010 Principal Investigators (2014) estimate, about 1 in 68 children in the United States is diagnosed with an ASD, consider the range of parenting challenges numerous parents now face. Cognitive Development Mother-child dyads from the Massachusetts site of the NICHD Study of Early Child Care and Youth Development participated in a home-based instructional task through which their scaffolding effectiveness was assessed. Cognitive characteristics of both mothers and children, as well as dyadic characteristics from infancy, were examined as predictors of effective dyadic scaffolding when the children were in first grade. Children's early mental development predicted the effectiveness of parent scaffolding collaborations over and above children's concurrent cognitive capabilities (Mulvaney et al., 2006). Likewise, toddlers who demonstrate more effortful control at age 18 months evoke more frequent use of cognitive tactics in their mothers in a teaching task at 30 months, controlling for initial values, and their effortful control at 30 months predicts mothers' cognitive orientation at 42 months (Eisenberg et al., 2010). More generally, the effectiveness of much parenting (e.g., discipline that emphasizes communication and reasoning) depends on children accurately construing parents' communications (Grusec & Goodnow, 1994). Reciprocally, simple child inattentiveness alters observed parenting (Browne et al., 2012), and mothers of children with a cognitive impairment are characterized as less sensitive and are more likely to engage in directive behaviors (Atkinson et al., 1999; Fenning et al., 2007; Moran, Pederson, Pettit, & Krupka, 1992; Niccols & Feldman, 2006). Children with early developmental delay in cognitive functioning are especially at risk of showing increased negative emotionality and behavior problems; in turn, these behaviors are associated with greater parenting stress and more psychological distress in their parents (Baker et al., 2002; Baker et al., 2005). Families face additional challenges when rearing children with impairments related to education, adaptive behavior, discipline, and the potential need for long-term care (Hodapp, 2002). Temperament and Personality

There is growing correlational and experimental evidence to suggest that child temperament and behavior influence parental cognitions and practices (Bates & Pettit, 2014). When, as mentioned earlier, Anderson et al. (1986) paired conduct-disordered boys with mothers of conduct-disordered boys and with mothers of normal boys, conduct-disordered boys elicited negative parenting practices from both sets of mothers. Clearly, in this experimental setting, which controls for genetic effects, temperament and personality characteristics of the child contributed to the parenting displayed. Across the infant's first year, evidence for bidirectional relations between infant temperament and parenting emerge. Infants who score low on a smiling and laughing temperament scale from ages 4 to 12 months have mothers who use more negative parenting (more mother-reported lax, overreactive, and overtalkative discipline) at 18 months, controlling for earlier mother personality (Bridgett, Laake, Gartstein, & Dorn, 2013), and having a temperamentally easy child or perceiving a child as temperamentally easy (relatively happy, predictable, soothable, and sociable) enhances mothers' feelings of competence and efficacy (Cutrona & Troutman, 1986; Dixon & Smith, 2003; Porter & Hsu, 2003; Putnam et al., 2002) and promotes warmer and more responsive parenting, whereas challenging children evoke parenting stress and harsh parenting (Balge & Milner, 2000; Deater-Deckard & Dodge, 1997; Gershoff, 2002; Mammen, Kolko, & Pilkonis, 2002; Rodriguez & Green, 1997). Difficult temperament reflects a child's tendency to react to stressors with negative emotions, including anger, irritability, fear, or sadness (Mertesacker et al., 2004; Rothbart, 2011). Early infant difficultness is associated with more parenting stress and less supportive coparenting across time. Paulussen-Hoogeboom, Stams, Hermanns, and Peetsma's (2007) meta-analysis revealed that child negative emotionality tends generally to be associated with lower parental warmth (in infancy and for high-SES families, negative emotionality may be associated with higher levels of parental warmth; Bates, Olson, Pettit, & Bayles, 1982; Crockenberg, 1986). One study (with autoregressive controls) found that high levels of infant distress at 1 year predicted reductions in mother supportive parenting from 1 to 2 years (Scaramella, Neppl, Ontai, & Conger, 2008). Children with more difficult temperaments often require more external, parental support and usually are recipients of less sensitive parenting (Belsky, 1984; Ciciolla et al., 2012). More generally, child temperament and behavioral styles that challenge parents may elicit harsher, less sensitive parenting behavior (Mills-Koonce, Propper, & AlYagon, 2007; Neppl et al., 2009; van den Boom & Hoeksma, 1994), although it is possible (albeit controversial) that negative emotionality also promotes more sensitive parenting in some situations (Leerkes & Crockenberg, 2002). Unsociable infant temperament predicts diminished maternal self-esteem (Farrow & Blissett, 2007), and child negative affectivity and conduct problems predict maternal reports of their own parenting as well as observations of parenting (Browne et al., 2012; Jenkins et al., 2003). Not unexpectedly, parents become warmer and less hostile in response to positive child and adolescent development and less warm in response to problematic child and adolescent functioning (L. R. Williams & Steinberg, 2011). Children's fearfulness and inhibition predict more maternal warmth, even controlling for earlier parenting (Lengua, 2006; Lengua & Kovacs, 2005). Parents' perceptions of child shyness at age 2 predict lower levels of parents' self-reported encouragement of child independence at age 4, even controlling for initial levels of encouragement of independence

and stability in child shyness (K. H. Rubin, Nelson, Hastings, & Asendorpf, 1999). High child social wariness at 18 months predicts lower levels of structured parenting at 27 months, even controlling for parenting at 18 months and for changes in child social wariness (Natsuaki et al., 2013); this particular study used adoptive children and their parents, thereby ruling out a shared-genes explanation for the child temperament effect. Children's developing personality continues to exert effects on parents. Using cohort-sequential latent growth curve modeling, Van den Akker, Deković, Asscher, and Prinzie (2014) studied mean-level personality development in children from 6 to 20 years of age and longitudinal bidirectional associations between child personality and maternal overreactive and warm parenting. In their 5-wave study, child personality predicted changes in both aspects of parenting. As every parent knows (too well), children's social successes lighten parents' burdens and hearten them, just as children's social difficulties stress and distress parents. Parents worry about their adolescents' participation in delinquent activities (Pasley & Gecas, 1984). If adolescents turn to deviant peers and increase their involvement in antisocial activity, parents often disengage (Dishion, Nelson, & Bullock, 2004; Dishion, Poulin, & Medici Skaggs, 2000). Thus, in response to children's antisocial behavior, negative parental control increases, and positive parent behaviors decline (Albrecht, Galambos, & Jansson, 2007; Ge et al., 1996; Reitz, Dekovic, & Meijer, 2006; Rueter & Conger, 1998; Stice & Barrera, 1995). Adolescents' antisocial behavior disrupts effective parenting, and parents become hostile and rejecting in the face of adolescent delinquency (Conger & Ge, 1999; Conger & Simons, 1997). Some authorities have argued that the impact of adolescent deviance on parenting is stronger than that of parenting on deviance (Albrecht et al., 2007; Jang & Smith, 1997; Kerr & Stattin, 2003). In brief, child temperament plays a pervasive role in family processes (E. F. Davis, SchoppeSullivan, Mangelsdorf, & Brown, 2009). Behavior Children's behaviors are important for adults' evaluations of children, and children's behaviors affect their parents in all sorts of ways. For example, mothers' disciplinary responses depend on the nature of the child's transgression more than their belief about how to discipline the child (Grusec & Kuczynski, 1980). Parents readily change their child-rearing practices in the face of children's behaviors (Holden, Thompson, Zambarano, & Marshall, 1997). From the moment of birth, certain baby signals effectively influence parenting: Infant cries motivate adults to approach and soothe (Ainsworth, Blehar, Waters, & Wall, 1978; Del Vecchio, Walter, & O'Leary, 2009; Konner, 2010; Soltis, 2004; Stallings, Fleming, Corter, Worthman, & Steiner, 2001), and their smiles encourage and reward adult proximity (Eibl-Eibesfeldt, 1989; Konner, 1991). Indeed, as to be expected in an alloparental species such as ours (Hrdy, 2009), child stimuli that activate parental brain systems in mothers (Rilling, 2013) activate many of the same parental brain systems in fathers (Kuo, Carp, Light, & Grewen, 2012; Wittfoth-Schardt et al., 2012) and nonparents (Caria et al., 2012; Glocker et al., 2009a). Infant distress and nondistress cues may prompt different schema that influence how parents feel and behave in the moment. Infant cries likely activate attachment-related schema and memories of the manner in

which parents' own emotional needs were or were not met in childhood, which in turn affect how aversive parents find crying, the manner in which they perceive infant cry signals, and their underlying beliefs about infants' emotions and how best to respond to them (Leerkes & Crockenberg, 2006; Leerkes, Parade, & Burney, 2010). The extent to which parent sensitivity to infant distress is distinct from sensitivity to nondistress or that both are dimensions of general sensitivity is important theoretically as well as practically. That is, if the two are distinct, they likely have different origins and may be related to different domains of child adjustment. And, if the two are distinct, intervention efforts could be developed to target the most relevant domain of parenting given the problem at hand (Leerkes, Weaver, & O'Brien, 2012). Hearing a baby cry decreases T in men when coupled with nurturant responses (van Anders, Tolman, & Volling, 2011), and aggressive behavior elicits harsh parenting (Anderson et al., 1986; Larzelere, 2000), whereas disclosures of information improves parental monitoring (Kerr & Stattin, 2000; Stattin & Kerr, 2000). Adolescents who disclose to their parents about their lives are more likely to have parents who refrain from reacting negatively (with sarcasm, judgment, or ridicule) to their spontaneous disclosures, and informed parents tend to be trusting parents (Kerr & Stattin, 2000; Kerr, Stattin, & Trost, 1999). Responsibility, independence, and freedom accompany child growth. The normal drive of the child and adolescent to establish her- or himself as a separate individual may clash with parents' desires to maintain dependence and inculcate their own values. Relationships Parent and child exist in a dyadic relationship, and the nature of their relationship has implications for parenting. As Maccoby (1992) long ago reflected, child rearing is an interpersonal activity that reflects the constant interplay and coordination of (at least) two individuals. Mother-child dyadic characteristics in infancy predict scaffolding when children are in first grade (Mulvaney et al., 2006). Stress derived from the mother-child relationship is associated with higher levels of neglect and affective withdrawal (Suchman & Luthar, 2001). The more rewarding their mutual interactions are, the more motivated are parents to seek quality interactions with their children again (Teti & Candelaria, 2002).

Summary Concurrent and predictive correspondences define the mutual influences that parent and child continuously exert on one another. The principle of transaction in development acknowledges that experiences shape the characteristics of the individual through time but that the characteristics of an individual shape his or her experiences (Bornstein, 2009; Sameroff, 2009). By dint of their unique characteristics and propensities—state of arousal, physical health, perceptual awareness, cognitive competence, emotional expressiveness, and individuality of temperament and personality—children actively contribute to their parenting and to producing their own development. Children influence which experiences they will be exposed to, and they interpret and appraise those experiences and so (in some degree)

determine how those experiences will affect them. Child and parent bring distinctive characteristics to, and each is believed to change as a result of, their interactions; both parent and child then enter the next round of interaction as changed individuals. For example, child temperament and maternal sensitivity operate in tandem to affect one another and eventually the later development (e.g., attachment status) of the child (Rothbart, 2011); parental psychological control and adolescent depressive symptoms exert reciprocal effects (Soenens, Luyckx, Vansteenkiste, Duriez, & Goossens, 2008); parents' use of corporal punishment predicts subsequent child behavior problems, while children with more behavior problems elicit more corporal punishment from their parents (Lansford et al., 2011); and mothers' and fathers' harsh verbal discipline provokes increases in adolescent conduct problems and depressive symptoms as adolescent misconduct elicits increases in mothers' and fathers' harsh verbal discipline (M. T. Wang & Kenny, 2014). Historically, it has been easy to assume that parents (and other adult socializers) are responsible for, say, gender-differentiated conduct in children. However, it is also the case that daughters elicit more feminine stereotypes and sons more masculine ones. Child effects on parents are in play and coexist with parent effects on children, and these mutual influences interact to consolidate socialization in children (Bornstein, 2013a). Another implicit assumption in parenting studies is that the effects of equivalent parenting are similar for all children. However, there is no reason to expect that a given parenting cognition or practice exerts the same effect on every child, and studies that pool parenting effects across children with different competencies, temperaments, and so forth might obscure child and parenting effects alike. A given parental cognition or practice likely has different effects on children with different temperaments. For example, Kochanska (1995, 1997) found that maternal responsiveness and the formation of a close emotional bond with the child fostered the development of conscience in bold, assertive children, whereas maternal gentle parenting techniques that deemphasized power assertion were more effective with shy, temperamentally fearful children. In making attributions about their behavior and in creating opportunities and environments (or in failing to do so), parents are often influenced by and respond to their children's characteristics (Bugental & Happaney, 2002; De Los Reyes & Kazdin, 2005). Furthermore, from early infancy children recognize their parents; show preferences for their sights, sounds, and smells; and over the course of just the first year of life, children develop deep and lifelong attachments to sensitive and responsive parents (Bornstein, Arterberry, & Lamb, 2014). In essence, then, children's parents receive a great deal “in kind” for their hard work and commitment: They are often recipients of unconditional love, and through their children they access immortality. In lieu of interpreting the link between warm parenting and adolescent adjustment to indicate an impact of parental warmth on adolescent well-being, it is equally reasonable to conclude that well-adjusted teens elicit warm parenting from their mothers and fathers (Kerr et al., 2003). Parents likely adjust to child behavior as much as they likely produce it. In the end, effects in socialization run in both directions—parent to child and child to parent— and this consistency is mutually reinforcing. Parents and their children co-construct

parenthood. Phonemes are the meaningful sound units of speech in a language, and they are constructed of subelements (so-called distinctive features). Jakobson (1941/1969) observed that the ontogenetic order of appearance of distinctive features in a Babel of languages including English, Swedish, and Japanese may be relatively fixed. Infants first produce front oral-cavity consonants (/m/, /p/) and mid- to back-vowels (/a/). Consonant-vowel combinations of these pairs are therefore frequently the earliest phonemes to appear in infant vocalizations. A well-known but nonetheless intriguing fact related to the ontogeny of vocalization is that two of these early and frequently appearing pairs, /ma/ and /pa/, tend to acquire particular meaning quite early in life. Murdock (1959) tallied the use of frontconsonant/back-vowel phonemes as parental descriptors in 1,072 languages and compared their frequency with other combinations (e.g., front-consonant/front-vowel, backconsonant/back-vowel, etc.) Fifty-seven percent of languages use the early-appearing frontconsonant/back-vowel class of infant sounds for initial parental kin terms. In short, parents around the world appear to adopt their infants' very first utterances as names for their own parenting role.

Determinants in the Context In addition to biological and psychological factors in themselves and their children, diverse contexts condition and channel cognitions and practices of parents. Context refers to “any event or condition outside the organism that affects or is affected by a person's development” (Bronfenbrenner & Crouter, 1983, p. 359). The contexts of parenting are complex, multidimensional, and structurally organized into levels that are linked with each other. Contexts influence parent characteristics, and parent characteristics influence their contexts. Rather than considering parent and context in isolation, there needs to be simultaneous consideration of both parent and context. Here, for convenience sake and heuristic purposes, I distinguish three kinds of contexts in which parenting is embedded and which together encompass contextual determinants of parenting. They are proximal contexts, social group contexts, and distal contexts. Proximal contexts include situation and demand, family structure and system, support networks, employment status, and neighborhood residence; social group contexts include SES, religion, ethnicity, and culture; and distal contexts include ecology, history, and even evolution. All these contexts encourage or discourage diverse patterns of parenting attitudes and actions, and so all are meaningful determinants to parenting.

Proximal Contexts Some prominent proximal context determinants of parenting include situations in which parents find themselves with children, tasks parents must accomplish, and immediate demands on parents as well as family structure and family systems, support networks that surround parents, their employment, and the neighborhoods where they live. Each is discussed in this section. Situation, Task, and Demand The supermarket checkout line is a more challenging context for the exercise of parental

authority than is the playground (Lecuyer-Maus, 2000). Although parenting cognitions and practices are relatively stable (in terms of rank-order consistency), both are flexible and can and do vary with situation, task, and demand (qualitatively and in terms of mean level; Bornstein et al., 2006; A. L. Miller et al., 2002). Thus, parental behavior is normally highly situation specific, and parents believe that they need to modify their attitudes and actions relative to specifics of their situation (e.g., Catron & Masters, 1993; Dix & Reinhold, 1991; Grusec & Kuczynski, 1980). Recall that Holden and Miller's (1999) meta-analysis showed that cross-situation parenting stability was lowest of three circumstances and change was greatest. Low-challenge situations (e.g., unstructured play) prompt one kind of parenting (P. J. Miller, Wang, Sandel, & Cho, 2002), as emotional and instrumental demands are relatively minimized (Ciciolla et al., 2012), but the degree to which parents are challenged during interactions affects their practices, as parents must make additional effort to remain calm and regulated under more taxing conditions (A. L. Miller et al., 2002; P. J. Miller et al., 2002). For example, maternal sensitivity during free play is lower than during bathing (Joosen, Mesman, Bakermans-Kranenburg, & van IJzendoorn, 2012). Maas et al. (2013) examined effects of situational variables on mothers' interactions at home with their 6-month-olds. They found that levels of interactive practices of mothers (e.g., sensitivity, stimulation) varied systematically across situations of free play, face-to-face play, and diaper change: During free play, mothers showed the highest levels of stimulation toward their infants; face-to-face interactions evoked more positive responsiveness; and in the goal-oriented diaper change situation, mothers hardly stimulated their infants and showed less positive regard. Likewise, mothers of children with and without developmental delay show similar levels of maternal sensitivity during free play; however, during a challenging task, mothers of children with developmental delay display lower levels of maternal sensitivity (Ciciolla et al., 2012). Maternal sensitivity during early childhood is therefore influenced by both the parenting context and specific child characteristics. Developmental risk and situation significantly contribute to the expression of maternal sensitivity, but their contributions depend on demands. Chaotic (unpredictable, nonroutine, inconsistent) environments (often but not always associated with poverty) undermine parenting self-efficacy and involvement, generate marital and parental stress, and undermine support and parent-child relationships (Corpaci & Wachs, 2002; Evans, 2004, 2006; Evans, Gonnella, Marcynyszyn, Gentile, & Salpekar, 2005; Wachs & Camli, 1991) and so child health and development (directly and indirectly). Household chaos describes an environment that is high in noise and low in regularity and routines and has been associated with caregivers who are less likely to monitor children's activities and resort more often to inconsistent discipline and physical punishment (Wachs, 2005, April]). All these sorts of contextualization further support the view of parenting as multidimensional, modular, and specific. Few studies have attempted to manipulate parenting demands within parent-child interactions by specifically contrasting interaction contexts (S. E. Martin, Clements, & Crnic, 2002; MillsKoonce et al., 2007; Pianta, Sroufe, & Egeland, 1989). However, studies which have indicated that the demands on parenting interact with other contextual factors. Perhaps most revealing of demand effectiveness on parenting are experimental treatments or interventions in which

parents are assigned randomly to a treatment/intervention (versus a control) where beliefs and behaviors of parents change. Experiments can modify parenting cognitions and practices (Forgatch & DeGarmo, 1999). Belsky, Goode, and Most (1980) found that reinforcing mothers' didactic interactions with their young children during play in an experimental group resulted in increases in mothers' didactic interactions (and in higher exploratory play in children) compared to a control group. Van den Boom (1994) trained low-SES mothers to respond sensitively to their children and in this way modified mothers' negative responses to child irritability (and reduced avoidant attachment in distress-prone children). Recall that Slep and O'Leary (1998) demonstrated that mothers could be led to believe that their children misbehaved intentionally, and they subsequently responded to their children with more negative affectivity and harsher discipline than mothers who were led to believe that their children misbehaved accidentally. And Stein et al. (2012) primed mothers to worry or ruminate and thereby evoked more negative thoughts, higher thought recurrence, and more selffocus and so inhibited their responsiveness and vocal interaction. As pointed out earlier, intervention trials can be interpreted as experimental manipulations that test parenting effects (P. A. Cowan & Cowan, 2002; DeGarmo, Eddy, Reid, & Fetrow, 2009; Dishion et al., 2008; Lunkenheimer et al., 2008; van Doesum, Riksen-Walraven, Hosman, & Hoefnagels, 2008). For example, Forgatch, Patterson, and DeGarmo (2005) found that fidelity to a therapy intervention protocol predicted change in parents' practices (and children's behavior). Experiments and interventions show that situation, task, and demand alter parenting. Family Structure Parenting is influenced by family configuration broadly construed. For example, marital status and father presence in the home moderate maternal emotional relationships with young children (Bornstein, Putnick, et al., 2012). Single-parent families typically experience less optimal interactions than do families consisting of two biological parents (Loeber et al., 2000; Sawhill, 2014). Single parenthood (Marquardt, Blankenhorn, Lerman, Malone-Colón, & Bradford Wilcox, 2012; Weinraub, Horvath, & Gringlas, 2002) and parents' separation and divorce (Hetherington & Stanley-Hagan, 2002) disrupt parenting and upset, change, and complicate family processes, as do subsequent transitions to stepparent or foster parent status (Haugaard & Hazan, 2002; Hofferth & Anderson, 2003). Notable in this connection is the incidence of child abuse in the stepparent population (Daly & Wilson, 1988). Many other vicissitudes of life (e.g., economic change) alter family structure and reverberate in changed parenting (Duncan, Magnuson, & Votruba-Drzal, 2015; Ganong et al., 2015). A common yet still enthralling change in parenting occurs when a second baby is born into the family. As reviewed earlier, parents treat their second-born children in many ways differently than they treat their firstborn children (Sulloway, 1996), even when first- and laterborns show no differences in behavior, indicating that some maternal behaviors reflect family structure (parity), and not child effects (Belsky et al., 1984). As reviewed earlier as well, experience in parenting matters, and mothers are prone to rate their firstborns as more temperamentally difficult (Bates, 1987), perhaps because first-time mothers are less at ease with their children and tend to perceive them as more demanding. Multiple births per se tax parenting (Bornstein

& Ruddy, 1984). Parents of twins versus singletons show differences in mean levels of parenting: Parents of twins feel less effective as parents, show less concern about the health and safety of their infants, and are more likely to use hostile-reactive behaviors than parents of singletons (Boivin et al., 2005). Evolutionary thinking suggests that parental effort expended on each successive offspring will decrease, resulting in smaller and less nourished offspring, a pattern that has been found in many different species ranging from invertebrates to birds to mammals (Mendl, 1988; Trumbo, 1996). The quality-quantity trade-off is documented in human beings in preindustrial societies, where larger family size has been linked to poorer growth and survival outcomes of children (Strassman & Gillespie, 2002), and in Western societies, where larger family size is associated with less direct parental involvement in key child care activities (Lawson & Mace, 2009). A Canadian cohort study showed that family size predicts maternal reports of their own parenting (Jenkins et al., 2003). Family Systems In a sense, the family system is another example of a situation effect on parenting. Parenting does not occur in a vacuum. It can be influenced by the presence of others; for example, fathers are more demanding of their sons when observed with their wives compared with father-son dyads alone (Buhrmester, Camparo, Christensen, Gonzalez, & Hinshaw, 1992). Thus, parents behave one way when interacting one on one with a child but another when the whole family is together (Cox & Paley, 2003; Deal, Hagan, Bass, Hetherington, & Clingempeel, 1999). Considerations of context and parenting necessarily invoke a family systems perspective. In family systems theory, what transpires in parenting is governed not only by the characteristics of the individual parent or child but also by patterns of transaction between the parent and others (Bornstein & Sawyer, 2005; Cox & Paley, 2003). Parenting develops and expresses itself in a family system that functions as an organized whole, composed of interdependent elements or subsystems that include individual parents as well as relationships among parents and others. Each element or subsystem within the family both affects and is affected by other elements; a change in any one element of the system can lead to changes in others. How responsive a mother or father may be at any given moment, for example, is determined not only by that parent's characteristic warmth and a child's characteristic responsiveness but also by the patterns the two parents have created jointly and therefore come to expect in their relationship. Moreover, a full family systems approach examines parenting in the context of all relationships within the family as well as relations between the family and its many enveloping social and distal contexts (such as socioeconomic class and culture, discussed shortly). For example, mothers who report supportive relationships with “secondary parents” (grandparents and the like) are more competent and sensitively responsive to their children than mothers lacking such relationships (Grych, 2002). Fathers' education predicts the quality of mother-child engagements (Tamis-LeMonda, Shannon, Cabrera, & Lamb, 2004). Reciprocally, conflict between spouses may cause parents to miss the sometimes-subtle signals children use to communicate their needs. For their part, even 1-year-old children are less likely to socially

reference their maritally dissatisfied father for information or clarification in the face of stress or ambiguity than are peers of maritally satisfied fathers (Parke, 2002). Reflecting a family systems perspective, marital relationships affect the quality of parenting (Grych, 2002; Tamis-LeMonda & Cabrera, 2002), just as the way parents work together as a coparenting team has multiple far-reaching consequences for parenting (Teubert & Pinquart, 2010). Coparenting refers to ways that parents (or parental figures) relate to each other in the role of parent (McHale et al., 2002; McHale & Lindahl, 2011; Van Egeren & Hawkins, 2004). Coparenting comprises multiple interrelated components: agreement or disagreement on childrearing issues; support or undermining of the parental role; and joint management of family interactions (Feinberg, 2003). Mutual emotional support and validation, modeling and sharing parenting skills, and buffering marital conflict or dissatisfaction from spilling over into relationships with children constitute some of the ways coparenting functions to moderate parenting. A coparent who has a positive and supportive relationship with a child tends to diminish the likelihood that the second parent will behave in a hostile manner toward the child (Conger et al., 2012). Another illustration of the moderating effects of family systems on parenting occurs when mothering interprets and conditions fathering, as mothers frequently act as gatekeepers to children's fathers (and other caregivers), encouraging or discouraging others' involvement in child rearing (Allen & Hawkins, 1999; Cannon, Schoppe-Sullivan, Mangelsdorf, Brown, & Sokolowski, 2008). Marital dissatisfaction predicts maternal reports of their own positive and negative parenting (Jenkins et al., 2003). More generally, gatekeeping refers to how each parent regulates the other's interactions with the child (Fagan & Barnett, 2003; Rane & McBride, 2000). In certain circumstances, fathers may be more involved with their children, as when mothers are more satisfied with paternal caregiving and assess fathers to be more competent at caring. Many (but not all) paternal influences on children are indirectly mediated through the father's impact on the mother (A. J. Walker & McGraw, 2000). Thus, maternal encouragement is linked to augmented father involvement in child rearing (“gate opening”; Schoppe-Sullivan, Brown, Cannon, Mangelsdorf, & Sokolowski, 2008), and maternal discouragement with diminished father involvement in child rearing (“gate closing”; Meteyer & Perry-Jenkins, 2010; Trinder, 2008). Many additional aspects of marital relationships affect the quality of mother-child and father-child relationships (Gable, Crnic, & Belsky, 1994; Fincham, 1998; Tamis-LeMonda & Cabrera, 2002). Support Networks Beyond the nuclear family system, all families are also embedded in, influence, and are themselves affected by larger social systems. These social systems include formal and informal support networks. Integration or isolation from potential support networks facilitates or hampers parenting. Support refers to psychological and tangible resources available to individuals through their relationships with family, friends, neighbors, work associates, and others. Normally, helpful support networks have positive effects on parents. Support from family members and other groups is associated with a successful transition to

parenthood (Bird, Peterson, & Miller, 2002; Elek, Hudson, & Bouffard, 2003). Mothers in supportive relationships are less harried and less overwhelmed, have fewer competing demands on their time, and, as a consequence are more available to their children and are more competent and sensitively responsive to their children than are mothers lacking such relationships (Crnic & Greenberg, 1990; Grych, 2002). A generalized benefit of support could occur because social networks provide parents with regular positive experiences and so bolster overall well-being, because supports provide a sense of predictability and stability in one's life situation and recognition of self-worth and because supports buffer stress and mental health issues in parents (Bird et al., 2002). Members of support networks can also teach and encourage parents in more developmentally appropriate caregiving. Several significant questions have motivated work in area of parenting supports. One concerns sources of support. In the Civitas Initiative et al. (2000) national survey, 70% of mothers of children under 3 reported that they relied on their spouse and 66% on their mothers for support. As noted previously, marital status and father presence in the home are critical to the young mother's well-being (L. D. Brown, Goslin, & Feinberg, 2011) and account for the quality of maternal emotional relationships with children (Bornstein, Putnick, et al., 2012). Emotional and child care support from a spouse (if not from other family members) are associated with well-being, greater life satisfaction, and more positive affect and competence (DeLongis, Capreon, Holtzman, O'Brien, & Campbell, 2004). Levitt, Weber, and Clark (1986) confirmed the importance of support from her spouse for mothers' well-being in intact families and extended the impact of that support to variation in the infant-mother relationship. In fact, intimate support from husbands appears to have the most general positive consequences for maternal competence (Crnic, Greenberg, Ragozin, Robinson, & Basham, 1983). When their children were 1, 3, and 5 years of age, parents reported on the emotional supportiveness they received from the other parent as well as their relationship status. Latent growth curve models revealed that, for both mothers and fathers, supportiveness was high at birth and perceived supportiveness at 1 year predicted supportiveness over time (Howard & Brooks-Gunn, 2009). Friends and relatives constitute other major sources of information and support for new parents, especially those with little child-rearing experience. In the Civitas Initiative et al. (2000) survey, 54% of mothers reported that they relied frequently on their child's doctor/pediatrician, 25% on nurses, and 20% of child care providers for information and advice. Health care professionals are less immediately accessible, but comprise an important reserve for all ages and social classes of parent. They are consulted most often about emergent or specific medical problems (Hickson & Clayton, 2002; Hulbert, 2003). At the same time, parents report that many anticipatory guidance topics are not covered in well-child visits and that, even after such visits, they can still use more information (Sanghavi, 2005). When parents were asked whether they had any developmental, behavioral, or health concerns about their child, and whether they shared those concerns with a professional, 29% reported at least one serious concern, and more than 20% had failed to share that concern (Golan et al., 2008). To establish social support networks for information about child rearing, middle-class American parents typically consult professionals, books, and magazines as well as relatives and friends (Clarke-Stewart, 1998; Young, 1991). Mothers in other social classes and cultures vary from

this “high-brow” mode of data collection, however. Informal supports appear to be especially important for ethnic minority families (Burton & Jarrett, 2000). Both informal support systems (the extended family) and formal ones (schools, child care, parent education programs, and professionals) influence parenting (Cochran & Niego, 2002). A second question in this literature concerns the types of support and the relative importance of the objective amount of support available to parents versus parents' perceptions of support. Support may be emotional, instrumental, or informational (Crockenberg, 1988). Quantitatively, more child care support is associated with higher-quality face-to-face interactions on the part of mothers (L. Levine, García Coll, & Oh, 1985). The more support adoptive parents receive (and the higher their expectations of life with a child), the more satisfaction they feel after their child arrives. Greater perceived support (both global and parenting support in particular) has been linked to lower levels of parenting stress, more positive parenting and relationships with children (Bonds, Gondoli, Sturge-Apple, & Salem, 2002; Izzo, Weiss, Shanahan, & RodriguezBrown, 2000; Suzuki, Holloway, Yamamoto, & Mindnich, 2009), and fewer negative interactions with children (Ensor & Hughes, 2010). Hashima and Amato (1994) found in a nationally representative sample that mothers' perceived social support was negatively associated with punitive and other negative parenting behaviors. The majority of existing research links global perceptions of support to indicators of parenting competence and suggests that it may matter more that parents feel support of some sort from some source than the specific amount, type, or source of support. D. M. Lee and Colletta (1983, April]) studied support-parenting relations in adolescent parents with young children. Mothers who were satisfied with their support reported that they were more affectionate with their children, whereas mothers who were dissatisfied with their support reported more hostility, indifference, and rejection of their children. Mothers' social support moderates the effects of daily hassles of parenting (Crnic & Greenberg, 1990). Parenting researchers have identified multiple antecedents of parenting knowledge qua a support (Cochran, 1993; Goodnow & Collins, 1990). Clarke-Stewart (1998), for example, reported that books and magazine articles were a popular source of advice about general child development, and their most frequent users were first-time and middle-class parents (Deutsch, Ruble, Fleming, Brooks-Gunn, & Stangor, 1988). Friends and relatives constitute a major base of information for younger parents with little parenting experience (Belsky, Youngblade, & Pensky, 1989). Mothers' own knowledge tends to relate to their formal, objective experience— exposure to books and manuals, pediatricians, and courses about child rearing—whereas previous informal or subjective experience with children (through babysitting or professional services alone) has been found to correlate not at all or even negatively with parenting knowledge (Frankel & Roer-Bornstein, 1982; MacPhee, 1981). Today, advice on parenting from outside the family abounds. It begins well before the birth of a child in so-called preconception care whose goal is to reduce the risk of adverse effects for women, fetuses, and neonates by optimizing women's health and knowledge before planning and conceiving a pregnancy (American College of Obstetricians and Gynecologists, 2005; http://www.cdc.gov/preconception/index.html). And it can be found in professional compendia that provide comprehensive medical treatises of prenatal, perinatal, and postnatal

development, such as A Guide to Effective Care in Pregnancy and Childbirth (Enkin, Keirse, Chalmers, & Enkin, 2000) and The A to Z of Children's Health: A Parent's Guide from Birth to 10 Years (Friedman, Saunders, & Saunders, 2013); in now-classic how-to books, such as Dr. Spock's Baby and Child Care (Spock & Needlman, 2013), Your Baby and Child (Leach, 2012), and What to Expect When You're Expecting (Murkoff & Mazel, 2008); in practical guides, such as Baby Care Basics (Friedman & Saunders, 2015) and Teach Your Children Well (M. Levine, 2013); in evidence-based academic compilations, such as the Handbook of Parenting (Bornstein, 2002b) and Parenting: A Dynamic Perspective (Holden, 2009); as well as in innumerable popular periodicals that overflow magazine racks in supermarkets, airports, and bookstores. Internet, blogs, and other social media regularly inform parenting and allow parents to both give and receive support (Brady & Guerin, 2010; Doty, Dworkin, & Connell, 2012). Parents report using these platforms to share thoughts and ideas about parenting as well as normalize their experiences as a parent by trying to determine whether or not circumstances, symptoms, or behaviors they experience are typical and expectable (Drentea & Moren-Cross, 2005; Kirby, Edwards, & Hughes, 2008). Parents often end discussion board posts with “Is this normal?” to solicit other parents' experiences. Parents use the Internet primarily for information seeking (Nichols, Nixon, Pudney, & Jurvansuu, 2009; S. K. Walker, Dworkin, & Connell, 2011) and more specifically for gathering parenting information (Bernhardt & Felter, 2004; Blackburn & Read, 2005; Madge & O'Connor, 2006; Radey & Randolph, 2009; Tuffrey & Finlay, 2002; Warren, Allen, Okuyemi, Kvasny, & Hecht, 2010). Parents report using the Internet to help them improve how they spend time with children (26%), care for children's health (19%), and learn new things (73%; Allen & Rainie, 2002). When parents search for information online, they often seek information related to their child's normative development, specific medical diagnoses, or strategies for developmentally appropriate parenting (Bernhardt & Felter, 2004; Tuffrey & Finlay, 2002). Overall, parents report more confidence and a greater sense of empowerment in their role as a parent as a result of gathering information online (Madge & O'Connor, 2006). For example, many parents use social media to ascertain other how other parents arrive at medical decisions regarding their children (Brady & Guerin, 2010). Parents often use the Internet to find and communicate with other parents who are in situations similar to their own (Tuffrey & Finlay, 2002). Parents of children with ASD, for example, utilize e-mail to validate one another's concerns, help each other cope with uncertainty, and make sense of their experiences as a parent of a child with autism (Huws, Jones, & Ingledew, 2001). In addition to parent education, parents access and report positive effects from online prevention and intervention programs (Lansford & Bornstein, 2007). In an evaluation of a 6month computer-mediated social support online community targeting young, single parents (Dunham et al., 1998), 42 women accessed the online network over 16,000 times and provided social support to each other through online discussion fora. This online resource provided an environment for developing close personal relationships and creating a sense of community. Overall, single mothers who participated reported that they felt decreased parenting stress. In an experimental study of over 800 Singaporean parents with children under the age of 6,

parents who accessed an online program providing information about children's development and tips for interacting with children increased in knowledge of children's speech and communication skills, children's social skills, and children's intellectual development. Parents who participated in the online program also reported feeling more confident that they were achieving their goal of being a successful parent (Na & Chia, 2008). First-time fathers who participated in an online intervention also reported significantly higher parenting self-efficacy scores than fathers in a comparison group (Hudson, Campbell-Grossman, Elek, Fleck, & Shipman, 2003). This New Fathers Network showed promise for being a successful intervention for providing nursing care knowledge to first-time fathers during the transition to parenthood. Employment Status Most parents work outside the home (Repetti & Wang, 2014). The nature of the workforce has changed dramatically; for example, the proportion of U.S. American mothers participating in the labor force increased from approximately 1 in 3 in 1950 to almost 2 in 3 in 2010 (U.S. Department of Labor, 2011). How individuals balance their roles in life, such as parent and employee, reflects in their effectiveness in those roles (Perry-Jenkins, Repetti, & Crouter, 2000). As noted earlier, mothers' role balance has more impact on child development than mothers' work status per se (Lerner & Galambos, 1985, 1986). Mothers who are unhappy with their work roles are more rejecting of their children, for example (Stuckey et al., 1982; Yarrow, Scott, De Leeuw, & Heinig, 1962). However, work can be affirming and have positive consequences for parenting. For example, in a sample of Korean mothers with infants, employment (both for consistently working mothers as well as for those transitioning to employment) was associated with fewer depressive symptoms and higher self-esteem, which in turn contributed to a warmer and more responsive parenting style (J. Kim & Wickrama, 2013). Thus, research tends to support a positive association between the balance struck by positive features of employment and parenting. Here, coparenting again enters the picture as a way for working women and men to balance their job and home lives (Sandberg, 2013). Overwork and difficulties at work, however, seem to undermine good parenting. Parents tend to reduce their levels of social engagement and expression of emotion after more demanding days at work (Repetti & Wang, 2014). Video recordings of the daily reunions of employed mothers and their preschool children at a work-site daycare center over the course of a week revealed that mothers spoke less and were less emotionally engaged (e.g., there were fewer observations of caring and loving/warm behavior) following more stressful work days (Repetti & Wood, 1997). There is also evidence for short-term increases in irritability, displays of anger, and use of discipline with children after difficult days at work (Repetti, 1994). Elevated levels of negative job spillover are linked to increased use of harsh discipline in fathers (Atzaba-Poria & Pike, 2008). Fathers are less likely to know about their children's experiences and activities if they describe their jobs as highly demanding or as interfering with family life (Bumpus, Crouter, & McHale, 2006), and those who work long hours and carry high workloads have more conflictive and less positive relationships with their adolescents (Crouter, Bumpus, Head, & McHale, 2001). More negative work environments predict more

intrusive and negatively valenced parenting behavior (Costigan et al., 2003). Thus, parents' work demands impinge on family functioning (Hsueh & Yoshikawa, 2007), just as family responsibilities impinge on parents' work performance (Buehler & O'Brien, 2011). Moreover, the impact on children of one parent's work conditions also depends on how the other parent (or partner) accommodates (Barnett & Gareis, 2007). More stressing still among women and men alike, unemployment, employment difficulties, and low earnings erode parents' positive connections to their children (Bianchi & Milkie, 2010). Indeed, at a macrosystem level, fertility itself declines in times of rising unemployment (Sobotka, Skirbekk, & Phillipov, 2011; Yule, 1906; Silver, 1965; Macunovich, 1996; R. Lee, 1990). Black, Kolesnikova, Sanders, and Taylor (2013) assessed the magnitude of long-term effects on completed fertility during the 1970s coal boom that increased male workers' incomes in coal regions of the United States and realized that that coal boom increased the completed fertility of affected cohorts by 3% while incomes were permanently increased by 6%. Analyzing more than 140 million birth records from the U.S. Vital Statistics natality data for the period 1975 to 2010, Currie and Schwandt (2014) detected short- and long-run negative responses of fertility to changes in the unemployment rate: Cohorts of women who faced high unemployment during their early 20s postponed fertility and had fewer children. Neighborhoods Neighbors may share norms and expectations for parenting and a collective willingness to implement both, even in the absence of close personal ties (Leventhal et al., 2015). Neighborhoods affect parents and parenting in many different ways: Neighborhoods likely influence parenting, parenting may mediate neighborhood characteristics on children, and neighborhoods can moderate parenting effects. First, the character of their neighborhood matters for parents (Franco, Pottick, & Huang, 2010; Guterman, Lee, Taylor, & Rathouz, 2009; Molnar, Buka, Brennan, Holton, & Earls, 2003; Skinner et al., 2014). Parents who live in more disadvantaged neighborhoods experience greater depression, more stress, and worse physical health (Franco et al., 2010; Guterman et al., 2009), all of which are associated with less warm and consistent parenting and strained family relationships (Conger & Donnellan, 2007; Elder, Eccles, Ardelt, & Lord, 1995; Leventhal & Brooks-Gunn, 2005; Ludwig et al., 2012). Parents who move to middle-class neighborhoods, compared with parents who stay in low-income neighborhoods, monitor their children less stringently and are less restrictive in their control and discipline (Briggs, Popkin, & Goering, 2010). Parents who are overwhelmed and stressed by their neighborhood fail to buffer their children from the negative aspects of the neighborhood. However, neighborhoodSES associations with parenting are complex, and (as discussed later) neighborhood affluence can also undermine parenting. However, research also shows that neighborhood social processes are associated with parents' well-being, cognitions, and practices separate and apart from SES. Neighborhood social processes and parenting are nuanced and potentially moderated by structural features. For example, lower neighborhood social cohesion is associated with greater parental depression (Kohen, Leventhal, Dahinten, & McIntosh, 2008), but higher neighborhood informal social control appears to aggravate the negative association

between parenting strain and parents' feelings of personal control (Carpiano & Kimbro, 2012). Second, parenting cognitions mediate links between neighborhood characteristics (social cohesion; Kohen et al., 2008; or disorder; Mrug & Windle, 2009) and parenting practices as well as children's development. In this sense, parenting may serve as an indirect pathway for the effects of neighborhoods on parents themselves or on their children. For example, neighborhood quality predicts parental efficacy, parental efficacy predicts reported parental involvement and monitoring, and both predict academic and socioemotional adjustment of adolescents (Shumow & Lomax, 2002). Parents are less likely to allow their children to spend time outside in neighborhoods that are less safe, restricting their children's options for out-ofschool-time physical activity (Schreier & Chen, 2012). Parental victimization by community violence and life stressors (parental strain) predict depressive, anxious, and hostile symptoms (parental mental health), which predict parental knowledge of their children's activities and child disclosure (parenting practices; Borre & Kliewer, 2014). Last, parenting appears to matter differently in different contexts. That is, specific combinations of parenting and neighborhood attributes interact to influence children's development. Thus, neighborhoods and parenting may be risk or protective factors, or neighborhoods may be a risk factor and parenting a protective factor, or vice versa (Cicchetti, 2010; Simons et al., 2002; Werner, 1995). Parenting and neighborhoods sometimes work in the same direction, for better or for worse. For example, negative, ineffective, or uninvolved parenting may threaten children's development in the context of high-risk neighborhoods (Roche, Ensminger, & Cherlin, 2007), and lower monitoring and permissive parenting in less safe neighborhoods and those with less residential stability and more disorder are detrimental for children (Beyers, Bates, Pettit, & Dodge, 2003). Similarly, harsh and inconsistent parental discipline in disadvantaged neighborhoods are associated with conduct disorder in children (Brody et al., 2003). By contrast, parenting strengths (e.g., use of authoritative control strategies) in neighborhoods where authoritative parenting is the norm may benefit children (Fletcher, Darling, Steinberg, & Dornbusch, 1995; Simons, Simons, Burt, Brody, & Cutrona, 2005). However, parental warmth, monitoring, and engagement can act as buffers in the context of low-SES or high-crime or disordered neighborhoods (Dearing, 2004; Rankin & Quane, 2002; Roche & Leventhal, 2009), just as neighborhood disadvantages can overcome advantages of typically beneficial parenting (Simons et al., 2002) or ineffective parenting is less problematic in the context of supportive neighborhoods (Browning, Leventhal, & BrooksGunn, 2005). The “goodness of fit” between parenting cognitions and practices and neighborhood characteristics conditions parenting, and the match between the two may also depend on parents' individual attributes (Carpiano & Kimbro, 2012) or ethnicity (Dearing, 2004; Lamborn, Dornbusch, & Steinberg, 1996) or other factors. For example, in high-SES neighborhoods, authoritarian parenting is associated with earlier sexual initiation, but high parental authoritarian control is associated with delayed sexual initiation in youth from lowSES neighborhoods (Roche et al., 2005). The outcome matters as well to the meaning of parenting in context. Authoritarian parenting is detrimental for children's behavior in the context of disadvantaged neighborhoods (Roche et al., 2007; Roche, Ghazarian, Little, &

Leventhal, 2011), but it is also associated with favorable academic outcomes (Dearing, 2004). A survey queried 3,115 adults living in 13 different neighborhoods in Texas to rate the effectiveness of different discipline strategies. Neighborhood conditions were related to the endorsement of yelling and threatening, time out and withdrawal of privileges, and explaining to children why certain behaviors are unacceptable, but they were not related to the endorsement of physical discipline. However, these associations differed by ethnicity. Neighborhood cohesion and collective efficacy were associated with decreased endorsement of yelling/threatening for African Americans and U.S.-born Latin Americans, and decreased endorsement of time out and withdrawal of privileges for European Americans. Neighborhood physical disorder, social disorder, and fear were associated with increased endorsement of time out and withdrawal of privileges among European Americans but were not associated with the endorsement of these strategies among ethnic minorities. Thus, parents' attitudes are structured by community context, but the influence of neighborhood conditions on parenting attitudes may differ for different ethnic groups (O'Brien, Caughy, & Franzini, 2005). Parents in impoverished neighborhoods enforce a variety of rules regarding children's peer relationships presumably to avert their children from the adverse effects of risky peers (Leventhal et al., 2015). For example, parents have rules about where adolescents may meet and regulations constraining the amount of contact they may have with peers. Following a change of residence, parents report that they engage in consulting strategies, such as talking with their adolescents and encouraging them to participate in peer activities (Vernberg, Beery, Ewell, & Absender, 1993).

Social Group Contexts Parents everywhere presumably want physical health, academic achievement, social adjustment, and economic security for their children (however those goals may be instantiated). It is likely, then, that even people in different social classes, professing different religions, belonging to different ethnicities, and from different cultures profess some similar general characteristics and competencies about parenting and wish to see and to promote some similar general characteristics and competencies in their offspring. Despite strong emphases on cultural differences, many cross-cultural studies of parent cognitions and practices report widespread similarities. In examples of the former, Mesman et al. (2015) sampled 751 mothers in 26 cultural groups from 15 countries around the globe and found strong convergence between maternal beliefs about the ideal mother and attachment theory's description of the sensitive mother across groups; Tulviste, Mizera, De Geer, and Tryggvason (2007) reported that child-rearing goals of mothers of 4- to 6-year-olds from Estonia, Finland, and Sweden were similar in all highly valuing self-maximization; Tudge and colleagues (1999) learned in a comparative study of parental values and beliefs in Estonia, Russia, the United States, and South Korea that middle-class mothers in all cultures were more likely to stress the importance of self-direction than the working-class mothers; and Posada and colleagues (1995) discovered that maternal beliefs about ideal child behavior were similar across seven Western and non-Western countries and overlapped with attachment theory's notion of the secure-base phenomenon. African American, Dominican immigrant, and Mexican immigrant mothers in the

United States reported the qualities they deemed desirable or undesirable in children age 1, 14, and 24 months. Mothers spontaneously referred to a common set of qualities, including selfmaximization and connectedness; most mothers approved of desirable qualities, such as achievement, and disapproved of undesirable qualities, such as improper demeanor (Ng, Tamis-LeMonda, Godfrey, Hunter, & Yoshikawa, 2012). In the practices domain, Bornstein, Putnick, Cote, Haynes, and Suwalsky (2015) examined rates, interrelations, and contingencies of vocal interactions in 684 mothers and their 5½month-old infants in diverse communities in 11 countries (Argentina, Belgium, Brazil, Cameroon, France, Israel, Italy, Japan, Kenya, South Korea, and the United States). Although they found that rates of mothers' and infants' vocalizations varied widely across communities, mothers' vocalizations to infants were consistently contingent on the offset of the infants' nondistress vocalizing, as infants' vocalizations were contingent on the offset of their mothers' vocalizing, and maternal and infant contingencies were significantly correlated. Thus, some parents' cognitions about their parenting and their children and some parents' practices are likely universal. Parenting presents the prospect of contributing to the development of a new life, and if there are universal values in this world, it is probable that nurturing, and not abusing, children falls among them. Similarities in parenting cognitions and practices could reflect evolutionary or biological bases of caregiving, the historical convergence of parenting styles, or the increasing prevalence of monistic parenting patterns through migration or dissemination via mass media. Even where ultimate goals may be similar, however, social classes, religions, ethnicities, and cultures may differ in proximal ways to achieve them (Bornstein, 1995). Furthermore, the parenting beliefs and behaviors of one's own social group may seem natural but may actually be rather unusual when compared with those of other groups. Thus, SES, religion, ethnicity, and culture constitute social groupings each of which likely conditions parenting. That is, all social classes, religions, ethnicities, and cultures prescribe certain beliefs and behaviors in their members and proscribe others (Maccoby, 2000). For parents, some prescriptions and proscriptions, such as the requirement that parents nurture and protect their offspring, are (as suggested) essentially universal. Others, such as what kinds of emotions can be expressed in public, vary from one social group to another, and social group variation in beliefs and behaviors is always impressive. For example, parents in some cultures talk to babies and see them as interactive partners (Bornstein et al., 2015), whereas parents in other cultures think that it is senseless to talk to babies (Ochs, 1988). Different social groups possess parenting ideas, approach parenting tasks, and value parenting outcomes differently (Cote, Bornstein, Haynes, & Bakeman, 2008; Goodnow & Lawrence, 2015). For these reasons, parents from different social groups differ in their opinions about the significance of specific competencies for their children's successful adjustment, differ in the ages they expect children to reach different milestones or acquire various competencies, and so forth. Specific goals arise, in part, out of unique expectations of adult members of specific social groups. In turn, distinct belief systems provide parents with a framework for interpreting their children's behaviors, guiding their interactions with their children, and determining the activities and opportunities that they supply for their children's development. Here I emphasize parenting differences

across social group contexts. Three additional considerations are preliminary to this discussion of social group membership vis-à-vis parenting. In some respects, SES and ethnicity, and ethnicity and culture, overlap, but in other respects they differ. Here I distinguish each. The reader needs to recognize that much of the extant literature addressing ethnicity has confounded SES and vice versa. Moreover, one ethnicity or different ethnic groups may inhabit the same place, but different cultural groups normally populate different parts of the world. Although not an exclusive definition (Bornstein, 2006b; LeVine, 2003), research sometimes succumbs to the mistaken tendency to lump across SES and ethnicity or ethnicity and culture, but each is heterogeneous, and across social classes, religions, ethnicities, and cultures, different parenting beliefs, different parenting behaviors, and (not unexpectedly) different patterns of child development prevail. There follows a brief overview of some central considerations about social group membership determinants of parenting. In the context of the exosystem ring of Bronfenbrennerian pressures, the order of social groups cited is arbitrary. Socioeconomic Status SES comprises income, education, and occupation of householders and is broadly influential in parenting (Bornstein & Bradley, 2003). Although parents in different strata behave similarly in certain ways, SES orders a wide variety of cognitions and practices of parents toward children as it does the home environment (Bornstein, Hahn, Suwalsky, & Haynes, 2003; Bradley & Corwyn, 2002; Hoff, Laursen, & Tardif, 2002). SES-related variation in parenting is diverse, conditioning the likelihood that parents serve their children spinach, display warmth in child rearing, and read books on child care. For example, using data from the Infant Feeding Practices study, which tracked the diets of more than 1,500 infants until age 1, Wen, Kong, Eiden, Sharma, and Xie (2014) found considerable differences in the solid foods mothers from different socioeconomic classes fed babies. Specifically, less-educated mothers from poorer households resorted to diets high in sugar and fat, whereas more-educated mothers from more resourced households adhered to diets that more closely followed infant feeding guidelines.The immediate consequence of poor infant diets is early weight gain and stunted growth. In the longer-term, the Infant Feeding Practices study analyzed data for the same children at age 6 and reported that infant feeding patterns translated into similar childhood eating habits. Parental education is a key factor in SES (Bornstein, 2003; Bornstein, Hahn, Suwalsky, & Haynes, 2003). It is likely that lower education levels are tied to poorer awareness of proper nutrition and infant feeding practices. That is, if parents lack a proper understanding of what infants should eat and how important proper nutrition is for child health, they may not select healthy foods when less healthy ones are more convenient. When Duncan and Brooks-Gunn (1997) statistically controlled for household income and other demographic characteristics, they found higher levels of maternal education were associated with better cognitive and educational outcomes in children. More educated parents are more likely to turn to professionals (e.g., doctors, nurses) and immediate family (e.g., spouses, parents) for parenting information and advice than are less educated parents. Parents with fewer years of education

are more likely to consult extended family (e.g., grandparents, aunts/uncles) for advice than are more educated parents. Lower-SES mothers refer to books or other written materials less readily as sources of information about child rearing and child development, whereas middleSES mothers report that reading material is their primary source of information (Furstenberg, Brooks-Gunn, & Chase-Lansdale, 1989; Hofferth, 1987; Young, 1991), and they seek out and absorb expert advice about child development (Lightfoot & Valsiner, 1992). Parents with more education possess more formal knowledge about child development norms and theories and about parenting practices (Bornstein et al., 2013; Conrad, Gross, Fogg, & Ruchala, 1992; Palacios & Moreno, 1996; Parks & Smeriglio, 1986). They have more knowledge about factors that reduce the risks of illness, and they implement practices at home to improve health. More educated parents not only spend more time with their children (Guryan, Hurst, & Kearney, 2008) but use that time differently in that they read to their children more frequently (Raikes et al., 2006) and speak to their children drawing on more sophisticated language and literacy skills (Arriaga, Fenson, Cronan, & Pethick, 1998; Hoff, 2003; Rowe, Pan, & Ayoub, 2005). Less educated mothers tend to stress traditional child-rearing goals more than mothers with higher levels of education (Tulviste et al., 2007). Similar dynamics obtain in fathers: Less educated Estonian and Finnish fathers are more likely to name desired characteristics related to conformity than more educated fathers. Educational level also drives mothers' expectations of when children should reach developmental milestones (P. D. Williams, Williams, Lopez, & Tayko, 2000). Higher-SES mothers generally give earlier age estimates for children's attainment of developmental milestones than lower-SES mothers (Mansbach & Greenbaum, 1999; von der Lippe, 1999). When Filipino mothers of preschool-age children were asked to estimate the age at which their child would acquire a variety of cognitive, psychosocial, and perceptual-motor skills (P. D. Williams et al., 2000), mothers with higher educational attainment had earlier expectations for children's cognitive and psychosocial development (e.g., emotional maturity, independence). Education also influences parental values for children. People with no more than a high school education tend to put more emphasis on teaching children obedience, religious faith, and being well mannered, whereas college graduates tend to place higher priority on teaching children empathy, curiosity, tolerance, and persistence (Doherty, Funk, Kiley, & Weisel, 2014). In short, the positive characteristics of caregiver (especially maternal) education vis-à-vis parenting are profound and pervasive. In developing nations, maternal education, net of other family socioeconomic indicators, is positively related to mother's own health care utilization as well as the nutritional status and health care utilization for her children (Abuya, Ciera, & Kimani-Murage, 2012; Ahmed, Creanga, Gillespie, & Tsui, 2013; Burchi, 2012). Using UNICEF's Multiple Indicator Cluster Survey, Bornstein, Putnick, Bradley, Lansford, and Deater-Deckard (2015) explored relations among mother education, household resources, and infant growth in 117,881 families living in 39 low- and middle-income countries. Mother education led to improved infant growth through availability of household resources. Unfortunately, in many low- and middle-income countries around the globe, instructional capital in the form of maternal education is limited: In 39 low- and middle-income countries,

for example, the median years of education for women aged 25 and over was only 5.17 in 2010 (Barro & Lee, 2010). Conversely, low SES and poor education are risk factors in parenting (and children's development). Low SES adversely affects mothers' psychological functioning and is associated with resort to harsh or inconsistent disciplinary behaviors (Atzaba-Poria & Pike, 2008; Pinderhughes, Dodge, Bates, Pettit, & Zelli, 2000). Parents who enjoyed fewer years of formal education and those with lower income hold less favorable attitudes about parenting in general (e.g., Clément & Chamberland, 2009; Pinderhughes, Dodge, Bates, Pettit, & Zelli, 2000) and sensitivity in particular (Emmen et al., 2012), and show a lower quality of actual parenting behaviors (e.g., Mesman et al., 2012). Around the globe, low- compared to middle-SES parents typically provide children fewer opportunities for variety in daily stimulation, have less appropriate play materials on hand, and engage in less total stimulation (Bornstein & Putnick, 2012). Significantly, working-class mothers converse with their children less, and in systematically less sophisticated ways, than do middle-class mothers (Hart & Risley, 1995; Hoff, 2003). These social class differences in maternal speech to children are widespread: In Israel, for example, middle-class mothers talk, label, and ask “what” questions more often than do lower-class mothers (Ninio, 1980). Parents in higher socioeconomic strata also change more flexibly and more rapidly in response to changes in developmental theory than do parents in lower socioeconomic strata (Bronfenbrenner, 1958). Lower-SES parents believe they have less control over the outcome of their children's development than do higher-SES parents (Elder et al., 1995; Luster, Rhoades, & Hass, 1989). Kohn (1963, 1969, 1979) hypothesized that social class differences in the requirements and expectations of fathers' jobs were related to differences in fathers' goals and expectations for their children. Lareau (2003) illustrated that lower-income parents adopt a more “natural growth” approach to child development, where they do not attempt to fill leisure time with structured activities, as is more typical of middle- and higher-income parents. Overall, financial and social stresses sap the general well-being and health of parents and demand attention and emotional energy from them (Magnuson & Duncan, 2002). In McLoyd, Aikens, and Burton's (2006) analysis, stresses on impoverished parents stemming from day-today struggles to find the resources to pay for food and rent, and the stresses of trying to cope with living in crowded housing and deteriorated, dangerous neighborhoods, undermine parenting skills and contribute to disorganizing family life (Toldson & Lemmons, 2013). Economic stresses thus erode parenting skills (Conger, Ge, Elder, Lorenz, & Simons, 1994) and reduce parents' attentiveness, patience, and tolerance toward children (Crnic & Low, 2002). Stress on parents is associated with decreased sensitivity to child cues, more negative feelings toward children, and harsh parenting styles (Sevigny & Loutzenhiser, 2010). Parenting in the upper class is not immune from disadvantage, however (Luthar, 2003; Luthar & Barkin, 2012; Luthar & Latendresse, 2005; Racz, McMahon, & Luthar, 2011). Affluence is sometimes associated with parents' lower social support and higher alcohol use (Briggs et al., 2010; Chuang, Ennett, Bauman, & Foshee, 2005; Fauth, Leventhal, & Brooks-Gunn, 2008). As we see, the economics of the family is open to future study (Duncan & Magnuson, 2002). For

example, standard economic theory (Becker, 1960) asserts that children are “normal goods” and fertility increases with income. However, higher female wages also make children more expensive, because childbearing and rearing have costs in terms of income lost. Empirical research supports these opposing forces (Aaronson, Lange, & Mazumder, 2011; Black, Kolesnikova, Sanders, & Taylor, 2013; Brueckner & Schwandt, 2014; Schaller, 2012). Religion Religion plays an central role in shaping core values and beliefs regarding family life, and aspects of religion are likely formative influences in parenting (Bengston, 2013; Gaunt, 2008; Mahoney, 2005; Mahoney, Pargament, Tarakeshwar, & Swank, 2008; Wilcox, 2002). A 2012 Gallup global poll reported that nearly 60% of the world's population self-identified as “religious” (WIN-Gallup, 2012). Most U.S. Americans describe themselves as religious (Mahoney, 2010), and 37% of all those with a religious affiliation (60% of White evangelical Protestants) in the United States say that religious faith is among the most important qualities to teach children (Doherty et al., 2014). (By contrast, just 3% of those who are unaffiliated say religious faith is among the most important qualities to instill.) In the United States, too, those with religious affiliations (particularly White evangelical Protestants) are also more likely than the religiously unaffiliated to prioritize obedience and being well mannered and somewhat less likely to say creativity, curiosity, or tolerance is important. That said, religion is an understudied determinant of parenting. In actuality, scholars have distinguished religiousness and spirituality as related but separate forces in parenting and family life (King & Boyatzis, 2015). Religiousness tends to refer to the extent to which an individual has a relation with a particular institutionalized religion (and is measured by, e.g., attendance at religious services), whereas spirituality is conceptualized in terms of an individual's personal quest for understanding answers to ultimate questions about life (and is measured by, e.g., perceptions of transcendence). Questions about religiousness tend to predominate in the extant literature and are usually global (Bartkowski & Xu, 2010) and too frequently overlook personal meanings derived from religious involvement (Mahoney, 2010; Mahoney et al., 1999; Volling, Mahoney, & Rauer, 2009). Parents who are religious are likely to manifest their values and beliefs through everyday interactions with others, including family members. Indeed, if religion guides parenting, we would expect somewhat rigid and less flexible caregiving because fixed factors, such as a religious creed, and not others, such as a child's needs, would determine parenting. For example, Conservative Protestants subscribe to the belief that corporal punishment as a disciplinary practice is beneficial for young children's socialization (Ellison, Bartowski, & Segal,1996; Thompson & Miller, 1997). Stronger religiosity is related to lower convergence between the mothers' beliefs about the ideal mother and the profile of the highly sensitive mother (Emmen et al., 2012). With possible exceptions of extremists, parental religiosity is positively associated with more effective parenting, communication, closeness, support, monitoring, conflict, and peer acceptance and negatively associated with authoritarian parenting (Snider, Clements, & Vazsonyi, 2004). A meta-analysis of studies associating religion and families found that religion relates to important outcomes in marriage and

parenting, including higher global marital satisfaction, family cohesion and authoritative parenting and lower rates of interparental conflict and divorce, with subsequent benefits in children's self-regulation (Mahoney, Pargament, Swank, & Tarakeshwar, 2001). DeMaris et al. (2011) assessed mothers' and fathers' reports regarding religion and parenting with infants; religious conservatism in mothers was related to higher levels of infant care, but lower levels in fathers. According to Ho (1994), Confucian ethics emphasize obligation to others rather than individual rights and provide a foundation for parent-child relationships through the notion of filial piety. Among such filial precepts are obeying, honoring, and providing for the material and mental well-being of one's parents; performing the ceremonial duties of ancestral worship; taking care to avoid harm to one's body; ensuring the continuity of the family line; and in general conducting oneself so as to bring honor and not disgrace to the family name. These underlying Confucian values are thought to be reflected in many Asian parents' goals for what it means to be a good parent and what constitutes a good or virtuous child. Two common construals offer insights into how religion might affect parenting. Sanctification refers to the extent to which parents view God as evident in their relationships with other family members and so view their parenting as imbued with religious and spiritual meaning. Sanctification is nondenominational. Sanctification in parenting is one way religion is manifested in everyday interactions between parents and children (Mahoney et al., 1999). For example, sanctification is associated with constructive discipline practices and greater marital satisfaction and diminished discord and with decreased conflict with children (Mahoney et al., 1999; Volling et al., 2009). Murray-Swank, Mahoney, and Pargament (2006) evaluated mothers spanking in relation to their theological conservatism in interaction with their sanctification (how much they imbued their parent role with sacred and divine qualities and saw parenting as “God's work”). Mothers' conservatism and sanctification interacted in predicting spanking: Theologically conservative mothers who also viewed their parental role as sanctified, sacred, and holy were more likely (than other conservative mothers) to spank their children, whereas mothers who were theologically liberal and also viewed their role as sacred and holy were less likely (than other liberal mothers) to spank. Religious coping describes the ways that individuals utilize religion to work through life challenges (Pargament, Smith, Koenig, & Perez, 1998). It concerns specific beliefs about God's role in times of distress and accounts for variance in health and well-being outcomes above and beyond the influences of nonreligious coping (Pargament, 1997). In addition, religious coping mediates the relation between general religious orientation and outcomes of life events. This functional aspect of religion might explain how sensitive and involved parents are with their children via-à-vis measures of attendance or global religiosity. Research has rarely considered whether religion has unique associations with parenting after accounting for related constructs that also may influence parenting. For example, as indicated, individuals reporting greater religiosity also tend to report greater marital satisfaction (Mahoney et al., 2001) and greater warmth and affection toward children (Wilcox, 1998), all of which predict more sensitive and involved parenting. In an investigation to rule out thirdvariable explanations for associations between religion and parenting, measures of civic engagement or conventional behavior (e.g., voting) were not found to diminish the association

between church attendance and father involvement (Wilcox, 2002). Examining the roles of other potentially important constructs linked to parenting will be critical for understanding how religiousness and spirituality each specifically shapes parenting. Ethnicity Parenting is conditioned by ethnicity and varies in meaningful ways among people of various ethnicities (as already noted). For example, parenting values is an instance case. A national survey in the United States revealed that (69%) African Americans say it is especially important to teach children religious faith; by contrast, about half of European Americans (51%) and Latin Americans (54%) say teaching children religious faith is important. African Americans (70%) are also more likely than European Americans (57%) to place importance on teaching children to be obedient (Doherty et al., 2014). Parenting style is often assessed along two separate broad dimensions, warmth (responsiveness and child-centeredness) and control (demandingness), which combined produce four parenting types: authoritative, authoritarian, permissive, and disengaged (Baumrind 1967, 1978, 1991; Maccoby & Martin, 1983). These styles are hypothesized to contribute differentially to child identity formation and cognitive and moral development. An authoritative parenting style combines high levels of warmth with moderate to high levels of control. In middle-class European American families, it is associated with children's achievement of social competence and overall better adaptation. Parents who are supportive in this way are often identified with being emotionally and physically responsive to their children's autonomy and prepare children for a world they regard as validating, trusting, and accepting. Authoritarian parenting, by contrast, is composited of high levels of control and low levels of warmth, and among European Americans it is generally associated with poorer child developmental outcomes. However, different ethnic groups in the United States have different approaches to parenting and value different outcomes. Parental control is equated with loving concern and associated with perceptions of warmth in Asian and Latin, but not European, American families (Chao & Tseng, 2002; Harwood et al., 2002). In European American families, parental warmth protects against children developing externalizing problems, but in African American families, it is a risk factor (Lau et al., 2006; Lorber, O'Leary, & Smith Slep, 2011). African American parents hold different beliefs about the world because of racism, inequality, and residence in neighborhoods that are unsafe (Ogbu, 1981). Thus, supportive parenting and its effects may look very different in different ethnic communities (Brody & Flor, 1998; Ispa et al., 2004). Tamis-LeMonda and Kahana-Kalman (2009) interviewed low-income, urban, African Americans as well as Mexican, Dominican, and Chinese immigrant mothers in maternity wards hours after the births of their babies. Mothers' views were assessed using open-ended questions, and their responses were coded as relevant to four main categories: child development, parenting, family, and resources. Mothers from the four ethnic groups varied in how much they spoke about child development, family, and resources. Relative to the other groups, Chinese immigrant mothers talked more about child development; African American and Dominican immigrant mothers talked more about resources; and Mexican immigrant mothers talked more about family. Latin American and African American mothers read to their

children less frequently as compared to European American mothers (Yarosz & Barnett, 2001), and Spanish-speaking Latin American families have fewer children's books available in the home than their non–Latin American counterparts (Raikes et al., 2006). As discussed earlier, O'Brien et al. (2005) surveyed adults living in 13 different neighborhoods in Texas about their attitudes regarding child discipline. Neighborhood cohesion and collective efficacy were associated with decreased endorsement of yelling/threatening for African Americans and U.S.born Latin Americans and decreased endorsement of time out and withdrawal of privileges for European Americans. Neighborhood physical disorder, social disorder, and fear were associated with increased endorsement of time out and withdrawal of privileges among European Americans but were not associated with endorsement of these strategies among ethnic minorities. Minority Latin and African American parents do not differ from one another in parenting knowledge but do differ from European American parents (García Coll & Pachter, 2002). Harwood, Miller, and Irizarry (1995) found that European American mothers underscore the importance of values such as independence, assertiveness, and creativity, when asked to describe their ideal child, whereas Latina American mothers underscore the importance of obedience and respect for theirs. Ethnicity also plays a role in parental supervision. European American, Latin American, and African American parents all similarly encourage their adolescents to establish same-ethnicity friendships and caution their adolescents about the challenges of cross-ethnic romantic relationships (Dornbusch et al., 1987; Hamm, 2001). Latin American and African American parents are consulted less by adolescents than are European American parents (Bornstein, Jager, & Steinberg, 2012). Asian American parents, as compared to European American parents, attribute the effort the child directs toward schoolwork as most important contributor to the child's academic success (Chao & Tseng, 2002). Culture Like social class, religion, and ethnicity, culture pervasively influences who parents (Leinaweaver, 2014) as well as how parents view parenting and how they parent (Bornstein & Lansford, 2010). Culture defies ready definition, but most scholars agree that culture embraces patterns of beliefs and behaviors, acquired through socialization, that distinguish social groups (Boyd & Richerson, 2005). The majority of research in parenting refers to a Western, educated, industrialized, rich, and democratic cultural database (Bornstein, 2010b; Henrich et al., 2010). Nonetheless, from extant comparative studies, we have learned that culture influences parenting (and child development) from very early in life in terms of when and how parents care for children, the extent to which parents permit children freedom to explore, how nurturant or restrictive parents are, which behaviors parents emphasize, and so forth (Benedict, 1938; Selin & Stone, 2009; Whiting, 1981). For example, in some cultures, children are reared in extended families where care is provided by many relatives; in others, children and their mothers are isolated from almost all social contexts. In some cultures, fathers are treated as irrelevant social objects (“honored guests”); in others, fathers assume complex and continuing responsibilities for children (“house husbands”). Japan and the United States maintain reasonably similar levels of modernity and living standards, and both are child-centered

societies, but the two differ dramatically in terms of history as well as beliefs and behaviors with respect to parenting. Traditional Japanese mothers expect early mastery of emotional maturity, self-control, and social courtesy in their children, whereas American mothers expect early mastery of verbal competence and self-actualization in theirs. American mothers promote autonomy and organize social interactions with their children so as to foster physical and verbal assertiveness and independence, and they promote children's mastery of the external environment. Japanese mothers organize social interactions so as to consolidate and strengthen closeness and dependency within the mother-child dyad, and they tend to indulge young children (Bornstein, Cote, Haynes, Suwalsky, & Bakeman, 2012; Morelli & Rothbaum, 2007). Parents in different cultures can differ radically in what they value. Consider play, for example: Parents in some cultures believe that play provides important developmentpromoting experiences; parents in others see play primarily to amuse; and parents in still others do not include play in their job description (Bornstein, 2007; Bornstein & Putnick, 2015). Tamis-LeMonda, Wang, Koutsouvanou, and Albright (2002) investigated child-rearing values in mothers of preschoolers from Greece, Taiwan, and the United States using openended probes and an ordering task. They identified four broad categories of values: decency (values emphasizing character), proper demeanor (values emphasizing appropriate relatedness and behaviors), self-maximization (values emphasizing the development of self potential and individuality), and sociability/lovingness (values emphasizing affective and social dimensions of relatedness). Cultural differences in all value categories emerged, echoing the unique cultural frameworks of the three societies. Likewise, expected developmental timetables in new mothers in Australia of Australian versus Lebanese heritage differ, showing that culture shapes mothers' expectations of their children much more than other seemingly more immediate factors, such as experiences observing their own children, comparing them to other children, and advice from friends and experts (Goodnow, Cashmore, Cotton, & Knight, 1984). Indeed, culturally defined beliefs are so powerful that parents sometimes act on them as much as or more than on what their senses tell them about their own children. Parents in Samoa, for example, reportedly think of young children as having an angry and willful character, and, independent of what children might actually say, parents consensually report that their children's first word is tae, Samoan for “shit” (Ochs, 1988). Among European Americans, self-esteem is thought to be important to healthy development in children, but among Taiwanese, self-esteem is less important or even thought to create psychological vulnerabilities. When A. L. Miller et al. (2002) compared European American and Taiwanese mothers' beliefs about child rearing and self-esteem, unsurprisingly they found that nearly all European American mothers spontaneously invoked self-esteem and spoke about the importance of building children's self-esteem. In contrast, few Taiwanese mothers talked about “self respect–heart/mind” (a Chinese term that approximates self-esteem). Asian Indian child rearing emphasizes close family bonds with parents highly involved in their children's lives. Individuals of all ages are expected to sacrifice and to contribute to the honor and welfare of the family (Jambunathan & Counselman, 2002). Rites of passage are associated with the onset of biological puberty, but there is no corresponding change in adolescents' status, responsibilities, or autonomy in decision making in Asian Indian families (Dasgupta,

1998), and even adult children are expected to remain at home until marriage and follow parental advice when dating or making career choices (Jambunathan & Counselman, 2002). By contrast, European American families tend to be nuclear, and children are generally viewed as transitional members who by adolescence are expected to individuate and pursue their own interests. Indeed, adolescence as it is conceptualized in Western society is a foreign concept in traditional Asian Indian society. Thus, Asian Indian parents hold negative views of their adolescents' independent and autonomous behaviors, behaviors that European American parents are trying to foster. Harwood et al. (1995) found that both working and middle-class Puerto Rican mothers valued respectfulness and proper demeanor in their children, but U.S. mothers wanted their children when grown up to possess the qualities of self-maximization. Parents in Sweden stress independence and assertiveness more than parents in India (Ekstrand & Ekstrand, 1987). Even very basic parenting practices vary widely across cultures. Konner (1977) recorded variations in the frequencies of African Kalahari San, Guatemalan, and Bostonian caregivers' vocalization to infants; Caudill and Weinstein (1969; see also Ferguson, 1977) found that U.S. mothers talk more to their babies than do Japanese mothers; and Richman et al. (1988) reported that North American, Swedish, and Italian mothers vocalized at higher rates than Kenyan (Gusii) mothers. (Many observers have noted the relative dearth of speech among African caregivers, be they Gusii speakers in Kenya or native speakers of Sooninke or Pular from west Africa; Dixon, Tronick, Keefer, & Brazelton, 1981; Rabain-Jamin, 1994; Richman et al., 1988.) Furthermore, European American mothers in Lafayette, Indiana, speak in response to their infants' vocalizations more than mothers in Nagoya, Japan, do (Fogel, Toda, & Kawai, 1988); mothers in Boston, Massachusetts, are more verbally responsive to their infants than are Gusii mothers in west Kenya (Richman et al., 1992); and urban Berliner mothers show higher levels of contingent vocal responsiveness compared with Nso mothers in rural Cameroon (Keller, Kärtner, Borke, Yovski, & Kleis, 2005). One dynamic attests to the influence of culture on parenting, and consequently the malleability of parenting. When parents reared in one culture migrate to another, they eventually adopt the parenting cognitions and practices of the new culture (Bornstein & Bohr, 2011; Bornstein & Cote, 2003, 2004a, 2004b, 2007; Cote & Bornstein, 2003, 2005; Cote et al., 2015). Notably, these rarely happen simultaneously, but practices seem to migrate sooner and more readily than cognitions. Multicultural comparative studies of parenting beliefs and behaviors, and even more challenging longitudinal ones, are increasing in incidence and sophistication (see Bornstein et al., 2015a, b; Lansford et al., 2011, 2014; Putnick et al., 2012). For example, mothers from Argentina, Belgium, Israel, Italy, Japan, and the United States were asked about their attributions for parenting successes and failures—if being able to successfully comfort their child when the child cries was due to their parenting ability (e.g., “I am good at this”), effort (e.g., “I have tried hard”), mood (e.g., “I am in a good mood”), task difficulty (e.g., “This is easy to do”), or a child characteristic (e.g., “My child makes this easy to do”). Among many culturally differentiated patterns of findings that emerged, Japanese mothers were less likely than mothers from all other nations to attribute success to their own ability and more likely to

indicate that, when they were successful, it was because of the child's behavior (Bornstein et al., 1992). Bornstein, Putnick, and Lansford (2011) evaluated similarities and differences in mean levels and relative agreement between mothers' and fathers' attributions and attitudes in parenting in nine countries: China, Colombia, Italy, Jordan, Kenya, the Philippines, Sweden, Thailand, and the United States. Mothers and fathers reported their perceptions of causes of successes and failures in caregiving and their progressive versus authoritarian child-rearing attitudes. Country differences emerged in all attributions and attitudes that were examined. In diverse domains of development, parenting practices relate differently to children's adjustment depending on the broader cultural context. A study of associations between parental corporal punishment and children's adjustment in China, India, Italy, Kenya, the Philippines, and Thailand disclosed that more frequent corporal punishment was linked to higher levels of child aggression and anxiety in all six countries, but the connection was weaker in countries where the use of corporal punishment was more culturally normative (Lansford et al., 2005). Bornstein and his colleagues (2015) observed mothers and their infants in 11 societies worldwide under ecologically valid, natural, and unobtrusive conditions to study their usual routines in the familiar confines of their own homes. They videorecorded mother-baby dyads and then used mutually exclusive and exhaustive coding systems to comprehensively characterize frequency and duration of six maternal caregiving practice domains (nurture, physical, social, didactic, material, and language), and they then asked about cultural similarities and differences in base rates of parenting in the six caregiving domains. Mothers in different cultures differed in every domain assessed. Moreover, mothers in no one culture surpassed mothers in all others in their base rates of parenting across domains. Parents' goals for their own parenting and for their children arise, in part, out of their society's expectations of its adult members (LeVine, 2003; Rosenthal & Roer-Strier, 2001). For example, many Western societies encourage independence, self-reliance, and individual achievement among their children, whereas many Asian and Latin cultures encourage interdependence, cooperation, and collaboration among peers (Triandis, 1995). These general values are associated with differences in parenting goals (Harwood, 1992; Ogbu, 1981). Keller et al. (2004) analyzed culturally informed parenting styles during infancy, as related to the sociocultural orientations of independence and interdependence. Free-play situations between mothers and 3-month-old infants were videorecorded in five cultural communities (West Africa, India, Costa Rica, Greece, and Germany) that differ according to their sociocultural orientations. They analyzed videorecords using coding systems that focused on four parenting systems, including body contact, body stimulation, object stimulation, and faceto-face contact. Two styles of parenting (distal and proximal) emerged across cultures that could be related to the sociocultural orientations of independence and interdependence. Li, Fung, Bakeman, Rae, and Wei (2014) compared conversations between European American and Taiwanese mothers and children (6–10 years) about good and poor learning. European American mothers mentioned mental activities and positive affect more, whereas Taiwanese mothers mentioned learning virtues and negative affect more. Schulze, Harwood, and Schoelmerich (2001) found that mothers from European American and Puerto Rican backgrounds commonly valued independent socialization, but European American mothers

emphasized emotional autonomy whereas Puerto Rican mothers focused on instrumental independence. Suizzo and Cheng (2007) analyzed mothers' socialization goals by parsing attributes of relatedness into family obligation and relational intimacy; relational intimacy was valued more highly by European American than Taiwanese mothers. By contrast, Taiwanese mothers were more likely than European American mothers to value agency-status goals concerning power and achievement. Some children are relatively relaxed when confronted with an unfamiliar situation and display little distress, whereas other children react to novel objects and situations with anxiety, try to remain close to their mother, and do not easily interact with unfamiliar people. These latter actions indicate behavioral inhibition. X. Chen et al. (1998) unearthed cultural differences in mothers' attitudes toward behavioral inhibition in children. Chinese mothers, who traditionally value mutual interdependence, view behavioral inhibition in toddlers as a positive trait. Behavior inhibition is associated with maternal acceptance of the child and belief in encouraging the child's achievement. In contrast, for Canadian parents of European origin, who traditionally hold a more individualistic orientation, behavioral inhibition is negatively associated with maternal acceptance and encouragement of children's achievement. Among traditional Chinese families, children who display higher levels of behavioral inhibition have mothers who are less likely to believe that physical punishment is the best way to discipline and who are less likely to feel angry toward the child. However, maternal punishment orientation was positively correlated with behavioral inhibition in Canadian families; mothers whose children displayed higher levels of behavioral inhibition are more likely to believe that physical punishment is the best discipline strategy. In short, behavioral inhibition in children is associated with positive attitudes in Chinese mothers and negative attitudes in Canadian mothers. In disciplining children, parents often escalate to corporal punishment. Today, 77% of American men and 65% of U.S. American women endorse the statement that sometimes a child needs a “good, hard spanking” (Child Trends, 2013), and almost 50% of American parents indicate that they had spanked their 2- to 5-year-old child in the last month (MacKenzie, Nicklas, Brooks-Gunn, & Waldfogel, 2011). Sanction of corporal punishment is widespread world-wide: A study of 30,470 families with 2- to 4-year-old children in 24 low- and middleincome countries revealed that 29% of parents reported believing that using corporal punishment is necessary to rear a child properly, and 63% of parents reported that their child had been corporally punished in the last month (Lansford & Deater-Deckard, 2012). Parents across cultures engage in child socialization practices to cultivate values and behaviors that promote adaptation to the social, economic, and ecological conditions of their society (Quinn, 2005). However, there are cultural differences in which values are prioritized to define the optimal endpoint of child development (Greenfield, Keller, Fuligni & Maynard, 2003; LeVine, 2003). Parenting cultural knowledge is often conceptualized as naive theories about how children progress toward idealized social roles, what influences their development, how skills are acquired, and the roles of different adults and environments in the developmental process (Keller et al., 2006; Rosenthal & Roer-Strier, 2001). Many argue that culturally distinct parenting beliefs provide parents with a framework for interpreting their

children's behaviors, guiding parents' interactions with their children, and determining the activities and opportunities that parents are willing to supply for their children's development. Parenting is a response to, and is supported by cultural beliefs about, what parents should do to promote their children's development.

Distal Contexts Distal contexts for parenting operate through proximal familial mechanisms and include, for example, ecology, time and history, and evolution. Here, I touch briefly on each. Ecology The ecological circumstances in which living organisms adapt and function are thought to bias them toward investing in either a relatively small number of “high-quality” offspring or a relatively large number of “low-quality” offspring. In humans, monogamy and biparental care are associated with allocation of resources toward few high-quality offspring: children who tend to have better health, survival, and developmental outcomes (Ellis, Figueredo, Brumbach, & Schlomer, 2009). Challenging, dangerous, and/or unpredictable environments favor low parental effort, and low parental effort is empirically associated with harsh, neglectful, rejecting, and insensitive parenting (Belsky et al., 1991). In an analysis of mostly preindustrial societies, Quinlan (2007) found that ecological stress, in terms of conditions of warfare, famine, and high pathogens, was associated with mothers decreasing parental care and terminating breastfeeding at earlier ages (De Baca, Figueredo, & Ellis, 2012). In reviews of maternal decision making under resource scarcity, Bateson (1994) and Hrdy (1999) concur. Specific patterns of child rearing can be expected to adapt to specific settings and needs. Thus, geographic conditions can be expected to play a consequential role in many aspects of parenting (O. Davis, Haworth, Lewis, & Plomin, 2012). Different geographic regions have different climatic, vegetative, and living conditions, and these circumstances can be expected to contribute to parenting cognitions and practices (R. R. Dunn, Davies, Harris, & Gavin, 2010). Insofar as tropical areas are denser with pathogens that lead to infections and poor child growth and mortality, for example, parenting in those ecologies can be expected to cope with greater challenges with respect to children's health and illness. Along these same lines, father-present social systems, monogamous marriage, and wide birth spacing are all more likely among human hunter-gatherers inhabiting harsh ecologies where biparental care is substantial and important for offspring survival and reproductive success (O. Davis et al., 2012; Draper & Harpending, 1988; Geary, 2000; Marlowe, 2003). Even different adapted mazeways in proximate ecologies give rise to divergent parenting. For example, nomadic hunter-gatherer Aka are more likely to hold and feed young infants in close proximity than are settled Ngandu farmers who are more likely leave infants by themselves, even though these two traditional groups live close to one another in central Africa (Hewlett, Lamb, Shannon, Leyendecker, & Schoelmerich, 1998). Aka parents are thought to maintain closer proximity to infants because this nomadic group moves in search of food more frequently than do more sedentary farmer Ngandu.

Within small rural communities characterized by subsistence economy, face-to-face interactions in small stable social networks, and lower levels of formal education and technology, socialization favors rearing children to promote group solidarity to meet communal needs. The extent to which the severity of exposure to social risk is related to parenting in the first 2 years of life was studied in diverse samples of families in two rural poor regions of the United States. The geographic isolation of families moderated associations between risk and expressions of maternal sensitivity and warmth as well as language and learning activities (Burchinal, Vernons-Feagans, Cox, & Key Family Life Project Investigators, 2008). By contrast, in urban settings characterized by formal schooling and commercial economic conditions that yield more fleeting interactions with strangers, socialization favors independence and competition in educational and economic realms (Greenfield, 2009). When Bornstein and colleagues (2008, 2012) investigated mothers' emotional availability to their children in contrasting rural and urban ecologies in three countries, they found patterns of parenting intrusiveness consonant with this urban-rural distinction. Time and History The bioecological view of parenting encompasses multiple scales of time. Every parent realizes that the days are long but the years are short. As mentioned earlier, parents do not vary in themselves much from day to day—parenting cognitons and practices tend to be stable—and family values are especially conservative. For example, Bomhoff and Gu (2012) found that East Asians continue to endorse traditional and conservative values despite their residence in economically developed nations, including Japan, South Korea, and Taiwan. Time of day and time of year influence parenting. For example, parents are more likely to spank their children in the evening than other times of the day (Clifford, 1959; Holden, Coleman, & Schmidt, 1995), and parenting differs during summer and winter months, consistent with parents' work schedules (Crouter & McHale, 1993). Over longer periods, parents also change on accounts of personal development, development in the child, and historical development. The ratio of adult-directed speech to child-directed speech increases across just the first postpartum year (Bornstein & Tamis-LeMonda, 1990), and parents of younger children spend much more time in child care than parents of older children (U.S. Bureau of Labor Statistics, 2013). “Parents who are highly effective at one stage in the child's life [are] not necessarily as effective at another… Similar behaviors do not necessarily produce the same effects at successive stages in [a] child's life” (Baumrind, 1989, p. 189). Parenting, like children's behavior, develops to become more hierarchical, differentiated, and complex (Sroufe & Jacobvitz, 1989). It is important to recognize that child-rearing cognitions and practices evolve and change with longer scales of time (Bronfenbrenner, 1958; French, 2002). In classical Rome, the pater familias dictated the literal life and death of children; later if any question arose about child care, mothers' rights prevailed; today in the postdivorce binuclear family, joint custody is fast becoming preferred law and the norm. The attitudes and actions of parents at any one time in any one generation may differ from those characteristic of parents in another. Change over time

in the parent, the context, and the relation between the parent and context is expectable. Thus, parenting values change over time along with societal changes. In essence, child-rearing cognitions and practices depend on the current cultural context as well as contexts of past and future (Cole, 1996; Valsiner, 2000). Across cultures, there is a general drift toward greater individualism that emphasizes individual self-sufficiency. Unsurprisingly, in their selfrepresentations young and middle-aged adults use more agency attributes, whereas older adults use communion attributes (Diel, Owen & Youngblade, 2004). On account of high rates of child mortality, parents earlier in history may have cared for but resisted emotional investment in the very young (Dye & Smith, 1986), an orientation that appears to persist where especially dire circumstances reign (Scheper-Hughes, 1989). One historian theorized that parents have generally improved in their orientation to and treatment of children because parents have, through successive generations, improved in their ability to identify and empathize with the special qualities of early childhood (deMause, 1975). Society at large is also witnessing the emergence of striking permutations in parenthood and in the constellation of the family structure that have plunged the family generally, and parenthood specifically, into an agitated state of question, flux, and redefinition (Ganong et al., 2015; Solomon, 2012). Peruse the panel of graphs in Figures 5.1 through 5.12. They evidence startling cultural shifts and social changes in central aspects of parenting even in so short a time as the second half of the twentieth century. In an analysis of parenting change from just 1989 to 2010, Park, Coello, and Lau (2014) noted that, relative to earlier-generation parents, later birth cohort parents in a number of Eastern and Western counties now endorse child independence, imagination, and determination and are less likely to value hard work, responsibility, obedience, thrift, and religious faith. Social change toward economic development may be associated with deemphasizing traditional mores, hierarchy, and work ethic while orienting parents to child-centered parenting. This pattern of change is consistent with emerging evidence concerning shifts toward materialism and individualism as economic fortunes increase at familial and national levels (Park, Twenge, & Greenfield, 2013). Research regarding birth cohort and social class suggests that economic development and schooling are especially influential in cultural change (Kashima et al., 2009; Inglehart & Baker, 2000; LeVine, 2003). Rapid cohort effects on child socialization have been documented in China, where social reticence formerly valued in traditionally collectivist Chinese contexts no longer appears adaptive in urban centers that favor selection of social initiative (X. Chen, Cen, Li, & He, 2005).

Figure 5.1 Mothers with College Education. Source: D. Cohn & G. Livingston (2013), Record share of new mothers are college educated. Pew Research Center. Retrieved from http://www.pewsocialtrends.org/files/2013/05/fertilityeducation_final.pdf

Figure 5.2 Unmarried Mothers. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.3 Mothers Ages 18–64 with a Young Child at Home. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.4 Age of First-Time Mothers. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.5 Mothers' expected number of children. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.6 Births to Unmarried Mothers. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.7 Hours Spent Working Outside the Home per Week by Mothers. Source: A. Caumont & W. Wang 2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.8 Hours Spent on Housework per Week by Mothers. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.9 Mothers with Children Younger than 18 in the Labor Force. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.10 Households with the Mother as the Primary or Sole Source of Income. Source: A. Caumont & W. Wang (2014). 5 questions (and answers) about American moms today. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/05/09/5-questions-and-answers-about-american-moms-today

Figure 5.11 Stay-at-Home Mothers in Poverty. Source: D. Cohn & A. Caumont (2014) 7 key findings about stay-at-home moms. Pew Research Center. Retrieved from http://www.pewresearch.org/fact-tank/2014/04/08/7-key-findings-about-stay-at-home-moms

Figure 5.12 Percentage of Single-Parent Families with Children under 18. According to Greenfield (2009), worldwide social change associated with economic development is altering cultural values and socialization environments as parents adapt to promote the success of their offspring in modernizing society. Greenfield reasoned that, as an environment shifts toward urbanization and economic development, individualist socialization waxes and collectivist socialization wanes. This line of thought echoes Inglehart's (1997) concept of transition from premodern to postmodern societal values. These kinds of studies reveal the imprint of historical time on parents' priorities in child socialization. Evolution

This examination of determinants of parenting began with biological factors in the parent that are themselves presumptive of grander evolutionary origins. So we have come full circle and end where we began. Driven by evolutionary pressures related to survival and reproduction, parents feel compelled to care for their (biological) offspring (Dawkins, 1976). Evolutionary theory might therefore elucidate some causes of variation in parenting (De Baca et al., 2012). For example, genetic relatedness among family members is a strong predictor of parental quality and even grandparental investment (O'Connor, Dunn, Jenkins, & Rasbash, 2006). Consider too the next predictions based on evolutionary thinking. First, as indicated, in an evolutionary perspective, a central principle governing parental effort is genetic relatedness. Parenting is selected to discriminatively favor biological offspring and to inhibit expending effort on the offspring of unrelated individuals. Research suggests that parents tend be warmer and engage in more caregiving of their biological children compared to stepchildren (which anecdotal evidence from a host of fairy tales seems to confirm; Bettelheim, 1976). In a comprehensive study of adolescent family relationships in stepfamilies, Hetherington and colleagues (1999) learned that biological relatedness between parent and adolescent was a potent determinant of parenting. Low genetic relatedness between adults and children in a family is associated with a higher probability of neglect, abuse, and homicide of children. The child abuse literature is consistent that children in stepfamilies are at greater risk for neglect and abuse compared to families with two biological parents (Daly & Wilson, 1988). Second, evolutionary-based research, using animal models and humans, has demonstrated that major biobehavioral reorganizations occur in the parents' physiology and behavior during periods of bond formation that lead to heightened sensitivity to infant cues, prepare parents for the challenging tasks of infant care, and give rise to expressions of the species-specific practices critical for child growth and adaptation to their ecocultural setting. Nurturance is a prerequisite for children's survival and well-being; seeing a child's survival to reproductive age enhances parents' probability of passing on their genes (Bjorklund et al., 2002). According to life history theory, third, a fundamental trade-off faced by parents is between current and future reproduction. Effort put into reproducing now will use energy or resources that cannot be used or saved for future reproduction. Thus, the greater the parent's probability of breeding in the future, the more the parent can be expected to withhold resources from current reproduction (Ellis et al., 2009; Trivers, 1972). Because age is a determining factor limiting future reproduction in women, this evolutionary view predicts that older mothers will systematically invest more in their current offspring. Consistent with this expectation, young mothers are more likely to elect to abort fetuses (Finer & Henshaw, 2006) and to neglect, abuse, or even murder their young children (J. Brown, Cohen, Johnson, & Salzinger, 1998). Conversely, as mothers age, their levels of parental effort allocated toward offspring tend to increase. Older mothers tend to be more satisfied in their role as a parent and spend less time away from their children (Ragozin, Basham, Crnic, Greenberg, & Robinson, 1982). These effects are not explained by maternal education, parity, gestational age, or income. A common proximate explanation for the effects of parental age on parental effort presumes that compromised parenting represents dysfunction or immaturity in young parents (e.g., inadequate

prefrontal development, poor judgment and planning, high impulsivity). An evolutionary explanation assumes that women of all ages adaptively calibrate their levels of parental effort in relation to future reproductive opportunities. This adaptive calibration may be instantiated at a proximate level through variation in such factors as maternal sentiments and interest in competing in nonparental activities. Fourth, from an evolutionary perspective, the level of genetic relatedness between a man and his assumed children, including uncertainty about that relatedness, is a crucial factor in determining whether (and how much) fathers care for their children (Geary, 2000). The expenditure of limited resources in a genetically unrelated child entails large fitness costs, so natural selection can be expected to favor discrimination and differential parental solicitude by fathers (Daly & Wilson, 1988). Relevant cues to paternity uncertainty include past and current extra-pair copulation rates, the mother's own sexual history, levels of trust and commitment in the father-mother relationship, and perceived similarities between the father and his putative children. Cues that indicate high paternity confidence support greater paternal investment and reduce the likelihood of paternal violence (Burch & Gallup, 2000). For example, across foraging groups in a representative data set of over 186 cultures around the world, monogamous societies are characterized by greater paternal investment (De Baca et al., 2012). Articulating four types of general evolutionary questions (we can appropriate to understanding the determinants of cognitions and practices in parenting), Tinbergen (1963) contended that evolutionary explanations concern phylogeny (What is the evolutionary history of a cognition or practice, and how did it change across time and species?), adaptation (Why is the cognition or practice the way that it is, and what selective advantages does it confer, or did it confer, to the organism?), mechanism (What is the cognition or practice like, and how does it work?), and development (How did the cognition or practice come to be, and how does it change across the life span?). These “evolutionary” questions pose contemporary challenges to parenting science.

Summary Parenting is conditioned by multiple spheres of influence in which the parent is embedded, including proximal, social group, and distal contexts. As this chapter has shown, researchers know a lot, but still not nearly enough, about the determinants of parenting. Some key questions that arise from the foregoing considerations concern the further specification of multiple processes by which parents' cognitions and practices come about and the multiple moderators that condition them (Bornstein, 2015; Bornstein, Bradley, Lutfey, Mortimer, & Pennar, 2011). In addition, there is a host of arenas about whose effects on parenting we know embarrassingly little. Consider ideology. The public is politically polarized, and ideological disparities between liberals and conservatives presumably color the cognitions and practices parents feel are important vis-à-vis children. We know that among the never married, large majorities of women (70%) and men (62%) alike seek a potential mate who shares their ideas about rearing children and consider that characteristic as more important than finding a partner who shares their moral and religious beliefs, has similar educational accomplishments, or even comes

from the same ethnic background (W. Wang & Parker, 2014). Doherty et al. (2014) conducted a nationally representative panel survey of randomly selected U.S. adults living in households that asked about the importance of teaching 12 different qualities to children that included responsibility, hard work, manners, persistence, tolerance, and faith. Respondents were also asked to choose the three traits that they felt are most important. Across ideologies, responsibility was viewed as important to teach children: 93% say teaching children to be responsible is “especially important,” and 55% rate responsibility as very important. Large majorities in different ideological groups also think it is essential for children to learn independence, hard work, and good manners. However, 86% of consistent liberals say it is important to teach children empathy; far fewer consistent conservatives (55%) agree. Liberals also prioritize teaching children curiosity (82%) and creativity (85%), compared with 60% or so in other ideological groups. By contrast, consistent conservatives think it is especially important for children to be taught religious faith (81%), and 59% say it is one of the three most important of 12 top qualities, whereas among consistent liberals, just 26% rate the teaching of religious faith as especially important and only 11% regard it as among the most important child-rearing qualities.

Translational Implications Practically speaking, parenting has positives, such as intimacy, nurturance, and rewards, which we want to encourage, but parenting can also be encumbered with negatives, such as frustration, anger, and violence, which we want to avert. It is a sad fact of everyday life that parenting children does not always go well or right. Infanticide was practiced historically, and although it is rare today, it is not unknown (Hrdy, 1999). Short of infanticide, vulnerable young are too commonly the victims of parenting that is neglectful (parents forget children in locked cars on hot days, fail to immunize them, or abandon them to wither in orphanages) and abusive (parents bear drug-addicted children, discipline them harshly, or traffic them). Infant cries draw a parent's attention but also trigger shaken baby syndrome (C. Lee, Barr, Catherine, & Wicks, 2007) and assaults (Cavanagh, Dobash, & Dobash, 2007). Every year, child protection agencies in the United States alone receive 3 million referrals for child neglect and abuse involving an estimated 6 million children younger than 5. About 80% of the children in investigated neglect and abuse cases are not removed from the home, although about 80% of perpetrators are parents, the vast majority biological parents. According to the National Research Council (2013), neglect and abuse amount to an estimated $80 billion per year in direct costs of hospitalization, law enforcement, and child welfare and indirect costs of special education, juvenile and adult criminal justice, adult homelessness, and lost work productivity. Many factors increase the risk of child abuse, but parental depression, parental substance abuse, and whether the parents were themselves neglected or abused as children rank high. Although usually protected, children are still too often exposed to relatively hostile and emotionally negative climates in the home, and parents in the front seat of the car fail to recognize that children in the backseat overhear what they say. By the current standards of meta-analytic evidence, parents contribute to their children's internalizing and externalizing

behavioral problems (Fearon, Bakermans-Kranenburg, van IJzendoorn, Lapsley, & Roisman, 2010; Groh, Roisman, van IJzendoorn, Bakermans-Kranenburg, & Fearon, 2012). Moreover, many parents resort to harsh verbal and physical treatment and even corporal punishment of children (Gershoff, 2002; Lansford et al., 2005; Mejia, Miewer, & Williams, 2006). From Finland and Denmark to Kenya, child-rearing violence predicts later child externalizing (Peltonen, Ellonen, Larsen, & Helwig-Larsen, 2010; A. T. Skinner, Oburu, Lansford, & Bacchini, 2014). The American Academy of Pediatrics (1998, p. 723) declared that “corporal punishment is of limited effectiveness and has potentially deleterious side effects,” recommending alternative determinants in that “parents be encouraged and assisted in the development of methods other than spanking for managing undesired behavior.” The American Academy of Pediatrics reaffirmed that pediatricians encourage parents to use alternate ageappropriate forms of nonaggressive discipline, such as distraction, time-outs, and offering explanations. The Convention on the Rights of the Child identifies violence toward children as a problem to be eliminated at an international level. The use of corporal punishment toward children has now been legally banned in all settings in 46 countries, and legislation is pending in several others (Global Initiative to End All Corporal Punishment of Children, 2013). Strong secular and historical trends operating in modern society—industrialization, urbanization, poverty, increasing population growth and density, and especially widespread dual-parental employment—constitute centrifugal forces on parenting. Because these societywide developments exert many unfortunately debilitative influences on parenthood, on parenting, and, consequently, on children and their development, a significant proportion of parents needs assistance to identify more effective strategies to optimize child care and to create more satisfying family relationships. Not surprisingly, a one-time U.S. commissioner of education preached that every child has a right to a “trained parent.” Parents are usually the most invested, consistent, and caring people in the lives of their children, so providing parents with skills and supports will help them respond more positively and effectively to social and child-related challenges. Professionals are normally a positive source of information related to parenting knowledge. As it is said, children do not come with a manual, and not all people naturally parent or know how to or can keep up with their constantly developing children. Parenting advice dates back at least to ancient Egypt, the Code of Hammurabi, and the pre-Socratics. In the Laws, Plato (ca. 355 BC) theorized about parenting. Many parents can use instruction. In 1914, the U.S. Children's Bureau published the first edition of the pamphlet “Infant Care,” which advised, among other things, that “Sunshine is as necessary for the baby as for the plant” (M. West, 1914, p. 10). Only a fraction of parents who could benefit from parent services receive them, however. Thus, organizations at all levels of society increasingly feel a mandate to intercede in child rearing and right social ills through parenting preventions and interventions. As a result, contemporary parenting has witnessed an explosive growth in information and support programs. One implication of the increasingly sophisticated view of the origins and conduct of parenting is that many determinants of parenting cognitions and practices are modifiable; thus, what we learn about parenting holds the promise of far-reaching practical implications. Slep and O'Leary (1998) found that mothers who were led to believe that their children intentionally

misbehaved responded to their children with greater negative affectivity and harsher discipline than mothers who were led to believe that their children accidentally misbehaved. As mentioned earlier, Stein et al. (2012) randomized mothers to a worry/rumination prime (WRP) or to a neutral prime (NP) and assessed mother-infant interactions before and after priming. Type of priming predicted maternal cognitions: The WRP resulted in more negative thoughts, higher thought recurrence, and more self-focus relative to the NP. Moreover, interaction effects between group and priming were significant: Compared with NP controls, the WRP inhibited maternal responsiveness and maternal vocal interactions. In the view of some critics, however, the trend to family professionalism also paradoxically leads away from a focus on parents as the proximal protectors, providers, and proponents of their progeny and toward entrusting child rearing to experts and the state. Nonetheless, contemporary parents from families along multiple risk continua might become more effective. Preventions and interventions designed to help parents come in a variety of venues (psychotherapy, classes, media), in a variety of settings (homes, schools, health clinics, houses of worship) and formats (individual, family, group), and with a variety of goals (from specific to universal). For example, some programs primarily target parents' cognitions and practices; others are focused on improving the quality of parent-child relationships. Competencies are knowledge, skills, abilities, personal characteristics, and attitudes, and competencies to adequately perform a task, duty, or role are (usually) learned (Roe, 2002). Competent parenting, crucial to child development and well-being, can be learned. Professionspecific competencies provide motivation and direction for learning as well as a means to judge the adequacy of parenting programs (Epstein & Hundert, 2002). Many experimental trials and parenting intervention programs have been devised. A review of 46 randomized trials provides evidence of prevention of a wide range of problem outcomes from 1 to 20 years later. However, this literature offers a paucity of evidence concerning processes that account for program effects (Sandler, Schoenfelder, Wolchik, & MacKinnon, 2011). Interventions to promote positive parenting have been touted to offer positive outcomes for children, but they often consume substantial resources and require rigorous appraisal. Biglan, Mrazek, Carnine, and Flay (2003) called for the integration of more science into societal parenting practices to prevent youth behavior problems and to promote “agreement on a set of consensus standards for selecting disseminable preventive interventions” (p. 438). A popular example is Triple P, the Positive Parenting Program, a multilevel behavioral family intervention (Prinz, Sanders, Shapiro, Whitaker, & Lutzker, 2009; Sanders & Kirby, 2010; Sanders, Markie-Dadds, & Turner, 2003). However, systematic reviews and meta-analyses comparing Triple P with waitlist or no-treatment comparison groups reveal only relatively small effect sizes on maternal and paternal outcomes (Wilson et al., 2012). Some interventions on determinants to improve parenting succeed: for instance, targeted interventions to develop maternal mentalizing and sensitivity and parents' positive practices (Juffer, Bakermans-Kranenburg, & van IJzendoorn, 2007; Lunkenheimer et al., 2008; Sadler et al., 2013). Others focus on factors that facilitate provider attitudes and skills (buy-in, support), staff selection and professional development (training, coaching), intervention and delivery (fidelity, accessibility), individual- and family-level factors (expectations, benefits), incentive

and retention strategies and those designed to overcome barriers in parent and program (costs, integration, caseload, competition). (See Axford, Lehtonen, Kaoukji, Tobin, & Berry, 2013; Damashek, Doughty, Ware, & Silovsky, 2011; Eisner & Meidert, 2011; Gottfredson et al., 2006; Korfmacher et al., 2008; Lees & Ronan, 2008; Mendez, Carpenter, LaForett, & Cohen, 2009; Mytton, Ingram, Manns, & Thomas, 2014; Shapiro, Prinz, & Sanders, 2012; Randolph, Fincham, & Radey, 2009; N. L. Weaver et al., 2008.) Unhappily, however, most interventions fail and do so for a wide variety of reasons, as when parents short-shrift intervention “homework” assignments, or when program staff fail to help parents access needed services, or when measurement is biased, or when the control group is contaminated through unintended diffusion of the parenting intervention. Some individuals or groups avoid seeking support out of embarrassment, stigma, or norms (in many communities, parents do not want to be told how to rear their children). (See Bodenmann, Cina, Ledermann, & Sanders, 2008; Fixen, Naoom, Blasé, Friedman, & Wallace, 2005; Goodson, Layzer, St. Pierre, Bernstein, & Lopez, 2000; Hebbeler & Gerlach-Downie, 2002; Okagaki & Bingham, 2005; Pinquart & Teubert, 2010; S. G. West, 2009; see Kaminski, Valle, Filene, & Boyle, 2008, for a meta-analysis of training programs.) To overcome these barriers and challenges, interventions need to adopt creative ways to market parenting and destigmatize the idea of parenting support (Chu, Farruggia, Sanders, & Ralph, 2012; Sanders, 2010). By deconstructing reasons for failures critically, it is possible to learn ways that future parenting interventions might succeed. Happily, Powell (2013) provides solid and timely guidance on central aspects of undertaking and evaluating parenting interventions, including preliminary work, sampling, research design, options for comparison conditions, measurement of implementation fidelity, and analysis of outcomes (Boutelle, Cafri, & Crow, 2011; DeGarmo & Forgatch, 2005; Flay et al., 2005; McCall & Green, 2004; Sandler et al., 2011; Shadish, Cook, & Campbell, 2002). To summarize some main points of that tutorial, parenting interventions should be based on theoretical and empirical models that specify the overarching goal of the intervention with regard to: anticipated improvements in specific parent outcomes; how change in malleable parenting variable(s) is targeted by the intervention; and how intervention components are expected to change focal parenting variable(s). Development of a coherent, measureable theory of change and articulation of a precise logic model are critical to an intervention functioning as intended. Randomly assigning parents or larger units in which parents are based to intervention and comparison conditions provides robust evidence of intervention effectiveness, assuming the outcome study is also well implemented. Intervention evaluations should routinely determine whether intervention effects are sustained beyond the intervention period (Olds et al., 2002). It can be guaranteed that intervention fidelity—the degree to which a program has been implemented as planned—is less than 100% ensured (Durlak & DuPre, 2008; Korfmacher, O'Brien, Hiatt, & Olds, 1999); but fidelity to an intervention protocol predicts change in parents' practices (Forgatch et al., 2005). Finally, no single intervention fits all (Barrera, Castro, & Steiker, 2011). So, for example, as an Institute of Medicine report (O'Connell, Boat, & Warner, 2009) recommended, interventions that successfully adopt and integrate cultural mazeways will experience likelier successes with both consumers and community agents (see Takanishi & Bogard, 2007).

As this chapter shows, individual factors are important determinants of parenting, and they are frequent targets of intervention. Higher reflective functioning is associated with sensitive parenting (Pajulo et al., 2009). Core concepts of reflective functioning and mindful parenting include, in addition to understanding mental states and intentions, full presence, nonjudgmental stances, emotional awareness, self-regulation, and compassion (Duncan, Coatsworth, & Greenberg, 2009). Certain mindfulness tools can help to successfully address pressing parenting requirements (Barth & Liggett-Creel, 2014; Kaminski, Valle, Filene, & Boyle, 2008; Sandler, Schoenfelder, Wolchik, & MacKinnon, 2011). Parents benefit from knowledge of how children develop. For example, of three cognitive measures, maternal education, IQ, and specific knowledge of child rearing and child development, knowledge uniquely explains age differences in mother-child emotional relationships (Bornstein, Putnick, et al., 2012). Therefore, the normative patterns and stages of children's physical, verbal, cognitive, emotional, and social development should be part of the knowledge base for parenthood. Parents need to know how to observe children and how to interpret and use what they learn. Informed child watching helps to clarify a child's level of development in relation to how parents want their children to behave and what parents want their children to learn and to accomplish. Observing also allows parents to identify potential trouble early and may help them respond to it more meaningfully. Parents need all manner of skills for managing children's behaviors. Knowledge of alternative methods of managing discipline and problem avoidance, for example, is basic. Parents need to understand the tremendous impact they have on their children's lives through the simplest things they do: their attention, expressed pleasure, listening, and interest. Parents need to know how to take advantage of everyday settings, routines, and activities to create learning and problem-solving opportunities that enhance childhood and parenthood. Parents need to be patient, flexible, and goal oriented—to call on their personal resources and extrapersonal sources of support—and they must command an ability to extract pleasure from their encounters with children. The responsibility for determining the child's best interests rests first and foremost with parents. Few ethical or sentient parents want to abrogate their child-rearing responsibilities (Baumrind & Thompson, 2002). Insofar as children's parents can be enlisted and empowered to provide children with experiences and environments that optimize their development, society is positioned to obviate after-the-fact remediation. In the dedication of Some Thoughts Concerning Education to Edward Clarke, of Chipley, Esq., the seventeenth-century moral philosopher John Locke (1692) wrote: “The well educating of their children is so much the duty and concern of parents, and the welfare and prosperity of the nation so much depends on it, that I would have every one lay it seriously to heart.” On these grounds, the doctrine of parental rights remains a fundamental premise to child rights. Parents sometimes do not know what to do, but they can find out; they sometimes do know what to do, but they still do not get

into the trenches and do it.

Conclusions Even the most vocal critics of parenting admit that “parents are the most important part of the child's environment and can determine, to a large extent, how the child turns out” (Harris, 1998, p. 15). We know a lot, but still not nearly enough, about what determines good or poor parenting. Parenting is multidimensional, modular, and specific and is also multilevel and multidetermined by parent, child, and contextual factors. This chapter has endeavored to overview the main determinants of parenting from that stance. To understand sources of variation in parenting, information about multiple domains is required, including biological and psychological characteristics of all individuals in the family, parents' families of origin, relationships between parents and between parents and children, and key individuals or institutions outside the family (including friends, peers, work, child care, ethnicity, culture, and historical era). Nuanced and complex principles of parenting, such as specificity, interdependence, transaction, and indirect and direct effects, do not make the study of determinants easy. Parenting sits at the confluence of many rivulets of influence; some flow from within the individual, whereas others spring from sources external to the parent in the child and in the multiple contexts that situate caregivers in the landscape parenting. Within complex developmental systems, it is unlikely that any single factor accounts for substantial amounts of variation in a criterion, and more multivariate and nuanced conceptualizations will explain parenting better than simpler ones with fewer variables. Parenting has many determinants, which is fortunate in the sense that in this way parenting is protected and the species continues, but parenting's multicausality is also demanding in the sense that many determinants require attention and a parent's job is never done.

References Aaronson, D., Lange, F., & Mazumder, B. (2011). Fertility transitions along the extensive and intensive margins (No. 2011–09). Working Paper, Federal Reserve Bank of Chicago. Abidin, R. R. (1992). The determinants of parenting behavior. Journal of Clinical Child Psychology, 21(4), 407–412. Abuya, B. A., Ciera, J., & Kimani-Murage, E. (2012). Effects of mothers' education on child's nutritional status in the slums of Nairobi. BMC Pediatrics, 12(1), 80. doi: 10.1186/1471– 2431–12–80 Adamson, L. B., & Bakeman, R. (1984). Mothers' communicative acts: Changes during infancy. Infant Behavior and Development, 7, 467–478. Adamsons, K., & Buehler, C. (2007). Mothering versus fathering versus parenting: Measurement equivalence in parenting measures. Parenting: Science and Practice, 7(3), 271– 303. doi: 10.1080/15295190701498686

Adler, A. (1928). Characteristics of the first, second, and third child. Children, 3(1), 14–52. Aguiar, M., & Hurst, E. (2007). Measuring trends in leisure: The allocation of time over five decades. Quarterly Journal of Economics, 122(3), 969–1006. doi: 10.3386/w12082 Ahmed, S., Creanga, A. A., Gillespie, D. G., & Tsui, A. O. (2013). Economic status, education, and empowerment: Implications for maternal health service utilization in developing countries. PLoS One, 5, e11190. doi: 10.1371/journal.pone.0011190 Ainsworth, M. D. S. (1969). Object relations, dependency, and attachment: A theoretical review of the infant-mother relationship. Child Development, 40(4), 969–1025. Ainsworth, M. D. S., Blehar, M. C., Waters, E., & Wall, S. (1978). Patterns of attachment: A psychological study of the strange situation. Hillsdale, NJ: Erlbaum. Albrecht, A. K., Galambos, N. L., & Jansson, S. M. (2007). Adolescents' internalizing and aggressive behaviors and perceptions of parents' psychological control: A panel study examining direction of effects. Journal of Youth and Adolescence, 36(5), 673–684. doi: 10.1007/s10964–007–9191–5 Allen, J. M., & Hawkins, A. J. (1999). Maternal gatekeeping: Mothers' beliefs and behaviors that inhibit greater father involvement in family work. Journal of Marriage and the Family, 61(1), 199–212. Allen, K., & Rainie, L. (2002). Parents online.Washington, DC: Pew Internet & American Life Project. Allik, J., & McCrae, R. R. (2004). Toward a geography of personality traits: Patterns of profiles across 36 cultures. Journal of Cross-Cultural Psychology, 35(1), 13–28. doi: 10.1177/0022022103260382 Altmann, J., Hausfater, G., & Altmann, S. A. (1988). Determinants of reproductive success in savannah baboons, Papio cynocephalus. In T. H. Clutton-Brock (Ed.), Reproductive success: Studies of individual variation in contrasting breeding systems (pp. 403–18). Chicago, IL: University of Chicago Press. Alvergne, A., Faurie, C., & Raymond, M. (2009). Variation in testosterone levels and malreproductive effort: Insight from a polygynous human population. Hormones and Behavior, 56(5), 491–497. doi: 10.1016/j.yhbeh.2009.07.013 Amato, P. R., & Booth, A. (2001). The legacy of parents' marital discord: Consequences for children's marital quality. Journal of Personality and Social Psychology, 81(4), 627–638. doi: 10.1037/0022–3514.81.4.627 American Academy of Pediatrics, Committee on Psychosocial Aspects of Child and Family Health (1998). Guidance for effective discipline. Pediatrics, 101(4), 723–728. American College of Obstetricians and Gynecologists. (2005, September). ACOG Committee

Opinion number 315. Obesity in pregnancy. Obstetrics and Gynecology, 106(3), 671. Anderson, K. E., Lytton, H., & Romney, D. M. (1986). Mothers' interactions with normal and conduct-disordered boys: Who affects whom? Developmental Psychology, 22(5), 604–609. Anjum, N., & Malik, F. (2010). Parenting practices in mothers of children with ADHD: Role of stress and behavioral problems in children. Pakistan Journal of Social and Clinical Psychology, 8, 18–38. Arber, S., & Timonen, V. (Eds.). (2012). Contemporary grandparenting. Bristol, UK: Policy Press. Armitage, C. J., & Conner, M. (2001). Efficacy of the theory of planned behaviour: A meta analytic review. British Journal of Social Psychology, 40(4), 471–499. Arriaga, R. I., Fenson, L., Cronan, T., & Pethick, S. J. (1998). Scores on the MacArthur Communicative Development Inventory of children from low- and middle-income families. Applied Psycholinguistics, 19, 209–223. Atkinson, L. (2012). Strategic decisions: Life history, interpersonal relations, intergenerational neurobiology, and ethics in parenting and development. Parenting: Science and Practice, 12(2–3), 185–191. doi: 10.1080/15295192.2012.683356 Atkinson, L., Chisholm, V. C., Scott, B., Goldberg, S., Vaughn, B. E., Blackwell, J.,… Tam, F. (1999). Maternal sensitivity, child functional level, and attachment in Down syndrome. Monographs of the Society for Research in Child Development, 64(3), 45–66. Atzaba-Poria, N., & Pike, A. (2008). Correlates of parenting for mothers and fathers from English and Indian backgrounds. Parenting: Science and Practice, 8, 17–40. doi: 10.1080/15295190701665698 Autism and Developmental Disabilities Monitoring Network Surveillance Year 2010 Principal Investigators. (2014). Prevalence of Autism Spectrum Disorder Among Children Aged 8 Years — Autism and Developmental Disabilities Monitoring Network, 11 Sites, United States, 2010. Centers for Disease Control and Prevention Morbidity and Mortality Weekly Report, 63 (2), 1–21. Avinun, R., Ebstein, R. P., & Knafo, A. (2012). Human maternal behaviour is associated with arginine vasopressin receptor 1A gene. Biology Letters, 8(5), 894–896. doi: 10.1098/rsbl.2012.0492 Axford, N., Lehtonen, M., Kaoukji, D., Tobin, K., & Berry V. (2013). Engaging parents in parenting programs: Lessons from research and practice. Children and Youth Services Review, 34(10), 2061–2071. doi: 10.1016/j.childyouth.2012.06.011 Bailey, J. A., Hill, K. G., Oesterle, S., & Hawkins, J. D. (2009). Parenting practices and problem behavior across three generations: Monitoring, harsh discipline, and drug use in the

intergenerational transmission of externalizing behavior. Developmental Psychology, 45, 1214–1226. doi: 10.1037/a0016129 Baker, B. L., Blacher, J., Crnic, K. A., & Edelbrock, C. (2002). Behavior problems and parenting stress in families of three-year-old children with and without developmental delays. Journal of Information, 107(6), 433–444. Baker, B. L., Blacher, J., & Olsson, M. B. (2005). Preschool children with and without developmental delay: Behaviour problems, parents' optimism and well-being. Journal of Intellectual Disability Research, 49(8), 575–590. doi: 10.1111/j.1365–2788.2005.00691.x Baker, J. K., & Crnic, K. A. (2005). The relation between mothers' reports of family-of-origin expressiveness and their emotion-related parenting. Parenting: Science and Practice, 5(4), 333–346. doi: 10.1207/s15327922par0504_2 Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2006). Gene-environment interaction of the dopamine D4 receptor (DRD4) and observed maternal insensitivity predicting externalizing behavior in preschoolers. Developmental Psychobiology, 48(5), 406–409. Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2008). Oxytocin receptor (OXTR) and serotonin transporter (5-HTT) genes associated with observed parenting. Social Cognitive and Affective Neuroscience, 3(2), 128–134. doi: 10.1093/scan/nsn004 Baldwin, A. L. (1946). Differences in parent behavior toward three- and nine-year-old children. Journal of Personality, 15, 143–165. Bales, K. L. (2014). Comparative and developmental perspectives on oxytocin and vasopressin. In M. Mikulincer & P. R. Shaver (Eds.), Mechanisms of social connection: From brain to group (pp. 15–31). Washington, DC: American Psychological Association. doi: 10.1037/14250–002 Balge, K. A., & Milner, J. S. (2000). Emotion recognition ability in mothers at high and low risk for child physical abuse. Child Abuse & Neglect, 24, 1289–1298. doi: 10.1016/S01452134(00)00188-5 Bandura, A. (1986). Social foundations of thought and action: A social cognitive theory. Englewood Cliffs, NJ: Prentice-Hall. Bandura, A. (1989). Human agency in social cognitive theory. American Psychologist, 44(9),1175–1184. Bandura, A. (1997). Self-efficacy: The exercise of control. New York, NY: Freeman. Barber, J. S., Axinn, W. G., & Thornton, A. (1999). Unwanted childbearing, health, and mother-child relationships. Journal of Health and Social Behavior, 40(3), 231–57. Barkley, R. A. (2012). Executive functions: What they are, how they work, and why they evolved. New York, NY: Guilford Press Press.

Barnard, K. E., & Solchany, J. E. (2002). Mothering. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 3–25). Mahwah, NJ: Erlbaum. Barnett, R. C., & Gareis, K. C. (2007). Shift work, parenting behaviors, and children's socioemotional well-being a within-family study. Journal of Family Issues, 28(6), 727–748. doi: 10.1177/0192513X06298737 Barrera, M., Jr., Castro, F. G., & Steiker, L. K. H. (2011). A critical analysis of approaches to the development of preventive interventions for subcultural groups. American Journal of Community Psychology, 48(3–4), 439–454. doi: 10.1007/s10464–010–9422-x Barrett, J., & Fleming, A. S. (2011). Annual research review: All mothers are not created equal: Neural and psychobiological perspectives on mothering and the importance of individual differences. Journal of Child Psychology and Psychiatry, 52(4), 368–397. doi: 10.1111/j.1469–7610.2010.02306.x Barro, R., & Lee, J. W. (2010). A new data set of education attainment in the world, 1950– 2010. Working Paper No. 15902. Cambridge, MA: National Bureau of Economic Research. Barry, C. T., Frick, P. J., & Grafeman, S. J. (2008). Child versus parent reports of parenting practices: Implications for the conceptualization of child behavioral and emotional problems. Assessment, 15, 294–303. doi: 10.1177/1073191107312212 Bartels, A., & Zeki, S. (2000). The neural basis of romantic love. NeuroReport, 11(17), 3829– 3834. Bartels, A., & Zeki, S. (2004). The neural correlates of maternal and romantic love. NeuroImage, 21(3), 1155–1166. doi: 10.1016/j.neuroimage.2003.11.003 Barth, R. P., & Liggett-Creel, K. (2014). Common components of parenting programs for children birth to eight years of age involved with child welfare services. Children and Youth Services Review, 40, 6–12. doi: 10.1016/j.childyouth.2014.02.004 Bartkowski, J. P., & Xu, X. (2010). Religion and family values reconsidered: Gender traditionalism among conservative Protestants. In C. G. Ellison and R. A. Hummer (Eds.), Religion, families, and health. Population-based research in the United States (pp. 106 – 125). New Brunswick, NJ: Rutgers Universtiy Press. Bates, J. E. (1987). Temperament in infancy. In J. D. Osofsky (Ed.), Handbook of infant development (2nd ed., pp. 1101–1149). New York, NY: Wiley. Bates, J. E., Olson, S. L., Pettit, G. S., & Bayles, K. (1982). Dimensions of individuality in the mother-infant relationship at six months of age. Child Development, 53, 446–461. Bates, J. E., & Pettit, G. S. (2014). Temperament, parenting, and social development. In J. Grusec & P. Hastings (Eds.), Handbook of socialization (2nd ed., pp. 372–397). New York,

NY: Guilford Press. Bates, J. E., Schermerhorn, A. C., & Petersen, I. T. (2012). Temperament and parenting in developmental perspective. In M. Zentner & R. C. Shiner (Eds.), Handbook of temperament (pp. 425–441). New York, NY: Guilford Press. Bateson, P. (1994). The dynamics of parent-offspring relationships in mammals. Trends in Ecology & Evolution, 9(10), 399–403. Baumrind, D. (1967). Child-care practices anteceding three patterns of preschool behavior. Genetic Psychology Monographs, 75, 43–88. Baumrind, D. (1978). Reciprocal rights and responsibilities in parent-child relations. Journal of Social Issues, 34, 179–196. Baumrind, D. (1989). Rearing competent children. In W. Damon (Ed.), Child development today and tomorrow (pp. 349–378). San Francisco, CA: Jossey-Bass. Baumrind, D. (1991). Effective parenting during the early adolescent transition. In P. A. Cowan & E. M. Hetherington (Eds.), Family transitions: Advances in family research series (pp. 111–163). Hillsdale, NJ: Erlbaum. Baumrind, D., & Thompson, R. A. (2002). The ethics of parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Practical issues in parenting (2nd ed., Vol. 5, pp. 3–34). Mahwah, NJ: Erlbaum. Becker, G. S. (1960). An economic analysis of fertility. In Universities-National Bureau (Ed.), Demographic and economic change in developed countries (pp. 209–240). New York, NY: Columbia University Press. Bell, R. Q. (1968). A reinterpretation of the direction of effects in studies of socialization. Psychological Review, 75(2), 81–95. Bell, R. Q. (1970). Sleep cycles and skin potential in newborns studied with a simplified observation and recording system. Psychophysiology, 6, 778–786. Bell, R. Q., & Harper, L. (1977). Child effects on adults. Hillsdale, NJ: Erlbaum. Bellinger, D. (1980). Consistency in the pattern of change in mothers' speech: Some discriminant analyses. Journal of Child Language, 7, 469–487. Belsky, J. (1984). The determinants of parenting: A process model. Child Development, 55(1), 83–96. Belsky, J., & Barends, N. (2002). Personality and parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social ecology of parenting (2nd ed., Vol. 3, pp. 415– 438). Mahwah, NJ: Erlbaum.

Belsky, J., Crnic, K., & Woodworth, S. (1995). Personality and parenting: Exploring the Mediating role of transient mood and daily hassles. Journal of Personality, 63, 905–929. Belsky, J., & de Haan, M. (2011). Annual research review: Parenting and children's brain development: The end of the beginning. Journal of Child Psychology and Psychiatry, 52(4), 409–428. doi: 10.1111/j.1469–7610.2010.02281.x Belsky, J., Fish, M., & Isabella, R. A. (1991). Continuity and discontinuity in infant negative and positive emotionality: Family antecedents and attachment consequences. Developmental Psychology, 27(3), 421–431. Belsky, J., Gilstrap, B., & Rovine, M. (1984). The Pennsylvania Infant and Family Development Project, Part 1: Stability and change in mother-infant and father-infant interaction in a family setting at one, three, and nine months. Child Development, 55(3), 692–705. Belsky, J., Goode, M. K., & Most, R. K. (1980). Maternal stimulation and infant exploratory competence: Cross-sectional, correlational, and experimental analyses. Child Development, 51(4), 1168–1178. Belsky, J., Taylor, D. G., & Rovine, M. (1984). The Pennsylvania infant and family development project, II: The development of reciprocal interaction in the mother-infant dyad. Child Development, 55, 706–717. Belsky, J., Youngblade, L., & Pensky, E. (1989). Childrearing history, marital quality, and maternal affect: Intergenerational transmission in a low-risk sample. Development and Psychopathology, 1, 291–304. Benasich, A. A., & Brooks-Gunn, J. (1996). Maternal attitudes and knowledge of childrearing: Associations with family and child outcomes. Child Development, 67(3), 1186–1205. Benedict, R. (1938). Continuities and discontinuities in cultural conditioning. Psychiatry: Journal for the Study of Interpersonal Processes, 2, 161–167. Bengston, V. (2013). Families and faith: How religion is passed down across generations. New York, NY: Oxford University Press. Berg-Nielsen, T., Vikan, A., & Dahl, A. A. (2002). Parenting related to child and parental psychopathology: A descriptive review of the literature. Clinical Child Psychology and Psychiatry, 7(4), 529–552. doi: 10.1177/1359104502007004006 Berlin, L. J., Brady-Smith, C., & Brooks-Gunn, J. (2002). Links between childbearing age and observed maternal behaviors with 14-month-olds in the Early Head Start Research and Evaluation Project. Infant Mental Health Journal, 23(1–2), 104–129. doi: 10.1002/imhj.10007 Bernhardt, J. M., & Felter, E. M. (2004). Online pediatric information seeking among mothers of young children: Results from a qualitative study using focus groups. Journal of Medical

Internet Research, 6(1), e7. doi: 10.2196/jmir.6.1.e7 Bettelheim, B. (1976). The uses of enchantment: The meaning and importance of fairy tales. . New York: Vintage. Bever, T. G. (Ed.). (1982). Regressions in mental development. Hillsdale, NJ: Erlbaum. Beyers, J. M., Bates, J. E., Pettit, G. S., & Dodge, K. A. (2003). Neighborhood structure, parenting processes, and the development of youths' externalizing behaviors: A multilevel analysis. American Journal of Community Psychology, 31(1–2), 35–53. doi: 10.1023/A:1023018502759 Bianchi, S. M. (2000). Maternal employment and time with children: Dramatic change or surprising continuity? Demography, 37(4), 401–414. doi: 10.1353/dem.2000.0001 Bianchi, S. M. & Milkie, M. A. (2010). Work and family research in the first decade of the 21st century. Journal of Marriage and Family, 72, 705–725. Biglan, A., Mrazek, P. J., Carnine, D., & Flay, B. R. (2003). The integration of research and practice in the prevention of youth behavior problems. American Psychologist, 58(6–7), 433– 440. doi: 10.1037/0003–066X.58.6–7.433 Bird, G. W., Peterson, R., & Miller, S. H. (2002). Factors associated with distress among support-seeking adoptive parents. Family Relations, 51, 215–220. doi: 10.1111/j.1741– 3729.2002.00215.x Biringen, Z., Matheny, A., Bretherton, I., Renouf, A., Sherman, M. (2000). Maternal representations of the self as parent: Connections with maternal sensitivity and maternal structuring. Attachment and Human Development, 2(2), 218–232. doi: 10.1080/14616730050085572 Bjoerklund, A., Lindahl, M., & Plug, E. (2006). The origins of intergenerational associations: Lessons from Swedish adoption data. Quarterly Journal of Economics, 121(3), 999–1028. Bjorklund, D. F., & Ellis, B. J. (2014). Children, childhood, and development in evolutionary perspective. Developmental Review, 34(3), 225–264. doi: 10.1016/j.dr.2014.05.005 Bjorklund, D. F., Yunger, J. L., & Pellegrini, A. D. (2002). The evolution of parenting and evolutionary approaches to childrearing. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 3–30). Mahwah, NJ: Erlbaum. Black, D., Kolesnikova, N., Sanders, S., & Taylor, L. (2013) Are children “normal”? Review of Economics and Statistics, 95(1), 21–33. Blackburn, C., & Read, J. (2005). Using the Internet? The experiences of parents of disabled children. Child: Care, Health and Development, 31(5), 507–515. doi: 10.1111/j.1365– 2214.2005.00541.x

Bodenmann, G., Cina, A., Ledermann, T., & Sanders, M. R. (2008). The efficacy of the Triple P-Positive Parenting Program in improving parenting and child behavior: A comparison with two other treatment conditions. Behavior Research and Therapy, 46, 411–427. doi: 10.1016/j.brat.2008.01.001 Bögels, S. M., & van Melick, M. (2004). The relationship between child-report, parent self report, and partner report of perceived parental rearing behaviors and anxiety in children and parents. Personality and Individual Difference, 37(8), 1583–1596. doi: 10.1016/j.paid.2004.02.014 Boivin, M., Pérusse, D., Dionne, G., Saysset, V., Zoccolillo, M., Tarabulsy, G. M.,… Tremblay, R. E. (2005). The genetic-environmental etiology of parents' perceptions and selfassessed behaviours toward their 5-month-old infants in a large twin and singleton sample. Journal of Child Psychology and Psychiatry, 46(6), 612–630. Bomhoff, E. J., & Gu, M. M. L. (2012). East Asia remains different: A comment on the index of “self-expression values,” by Inglehart and Welzel. Journal of Cross-Cultural Psychology, 43(3), 373–383. doi: 10.1177/0022022111435096 Bonds, D. D., Gondoli, D. M., Sturge-Apple, M. L., & Salem, L. N. (2002). Parenting stress as a mediator of the relation between parenting support and optimal parenting. Parenting: Science and Practice, 2(4), 409–435. doi: 10.1207/S15327922PAR0204_04 Bornstein, M. H. (1989). Sensitive periods in development: Structural characteristics and causal interpretations. Psychological Bulletin, 105, 179–197. Bornstein, M. H. (Ed.). (1991). Cultural approaches to parenting. Hillsdale, NJ: Erlbaum. Bornstein, M. H. (1995). Form and function: Implications for studies of culture and human development. Culture and Psychology, 1(1), 123–137. Bornstein, M. H. (Ed.). (2002a). Handbook of parenting (2nd ed., Vols. 1–5). Mahwah, NJ: Erlbaum. Bornstein, M. H. (2002b). Parenting infants. In M. H. Bornstein (Ed.), Handbook of parenting: Children and parenting (2nd ed., Vol. 1, pp. 3–43). Mahwah, NJ: Erlbaum. Bornstein, M. H. (2003). Maternal education. In J. R. Miller, R. M. Lerner, L. B. Schiamberg, & P. M. Anderson (Eds.), Human Ecology: An Encyclopedia of Children, Families, Communities, and Environments (Vol. 2, pp. 476–478). Santa Barbara, CA: ABC-CLIO. Bornstein, M. H. (2006a). Parenting science and practice. In K. A. Renninger & I. E. Sigel (Eds.), Child psychology in practice. In W. Damon & R. M. Lerner (Editors-in-Chief), Handbook of child psychology (6th ed., Vol. 4, pp. 893–949). Hoboken, NJ: Wiley. doi: 10.1080/15295192.2001.9681208 Bornstein, M. H. (2006b). Some metatheoretical issues in culture, parenting, and

developmental science. In Q. Jing, M. R. Rosenzweig, G. d'Ydewalle, H. Zhang, H. C. Chen, & K. Zhang (Eds.), Progress in psychological science around the world, Vol. 2: Social and applied issues (pp. 245–260). Hove, UK: Psychology Press. Bornstein, M. H. (2007). On the significance of social relationships in the development of children's earliest symbolic play: An ecological perspective. In A. Gönçü & S. Gaskins (Eds.), Play and development: Evolutionary, sociocultural, and functional perspectives (pp. 101– 129). Mahwah, NJ: Erlbaum. Bornstein, M. H. (2009). Toward a model of culture↔parent↔child transactions. In A. Sameroff (Ed.), The transactional model of development: How children and contexts shape each other (pp. 139–161). Washington, DC: American Psychological Association. doi: 10.1037/11877–008 Bornstein, M. H. (2010a). From measurement to meaning in caregiving and culture: Current challenges and future prospects. In C. M. Worthman, P. M. Plotsky, D. S. Schechter, & C. A. Cummings (Eds.), Formative Experiences: The Interaction of Caregiving, Culture, and Developmental Psychobiology (pp. 36–50). New York: Cambridge University Press. Bornstein, M. H. (Ed.). (2010b). The handbook of cultural developmental science. New York, NY: Psychology Press. Bornstein, M. H. (2012). Cultural approaches to parenting. Parenting: Science and Practice, 12, 212–221. Bornstein, M. H. (2013a). Mother-infant attunement: A multilevel approach via body, brain, and behavior. In M. Legerstee, D. W. Haley, & M. H. Bornstein (Eds.), The infant mind: Origins of the social brain (pp. 266–298). New York, NY: Guilford Press Press. Bornstein, M. H. (2013b). Parenting X Gender X Culture X Time. In W. B. Wilcox & K. K. Kline (Eds.), Gender and Parenthood: Biological and Social Scientific Perspectives (pp. 91–119). New York: Columbia University Press. Bornstein, M. H. (2013c). The specificity principle in parenting and child development: Everything in moderation. Unpublished manuscript, Eunice Kennedy Shriver National Institute of Child Health and Human Development, Bethesda, MD. Bornstein, M. H. (2014). Human infancy…and the rest of the lifespan. Annual Review of Psychology, 65, 121–158. Bornstein, M. H. (2015). Children's parents. In M. H. Bornstein & T. Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R. M. Lerner (Editor-inChief), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 55– 132). Hoboken, NJ: Wiley. Bornstein, M. H., Arterberry, M. E., & Lamb, M. E. (2014). Development in infancy: A contemporary introduction (5th ed.). New York, NY: Psychology Press.

Bornstein, M. H., Arterberry, M. E., & Mash, C. (2013). Differentiated brain activity in response to faces of “own” versus “unfamiliar” babies in primipara mothers: An electrophysiological study. Developmental Neuropsychology, 38, 365–385. doi: 10.1080/87565641.2013.804923 Bornstein, M. H., & Bohr, Y. (2011). Immigration, acculturation, and parenting. In M. H. Bornstein (Topic Chapter Ed.), Immigration – Family Transitions in R. E. Tremblay, M. Boivin, & R. deV. Peters (Eds.), Encyclopedia of Early Childhood Development. Ottawa, Canada: Center of Excellence for Early Childhood Development. Bornstein, M. H., & Bradley, R. H. (Eds.). (2003). Socioeconomic status, parenting, and child development. Mahwah, NJ: Erlbaum. Bornstein, M. H., Bradley, R. H., Lutfey, K., Mortimer, J. T., & Pennar, A. (2011). Contexts And contents of socialization: A life-span perspective. In K. Fingerman, C. Berg, T. Antonucci, & J. Smith (Eds.), Handbook of lifespan development (pp. 839–883). New York, NY: Springer. Bornstein, M. H., & Cote, L. R. (2003). Cultural and parenting cognitions in acculturating cultures: II. Patterns of prediction and structural coherence. Journal of Cross-Cultural Psychology, 34, 350–373. Bornstein, M. H., & Cote, L. R. (2004a). Mothers' parenting cognitions in cultures of origin, acculturating cultures, and cultures of destination. Child Development, 75, 221–235. Bornstein, M. H., & Cote. L. (2004b). Who is sitting across from me? Immigrant mothers' knowledge about children's development. Pediatrics, 114, 557–564. Bornstein, M. H., & Cote, L. C. (2007). Knowledge of child development and family interactions among immigrants to America: Perspectives from developmental science. In J. Lansford, K. Deater-Deckard, & M. H. Bornstein (Eds.), Immigrant Families in Contemporary Society (pp. 121–136). New York: Guilford. Bornstein, M. H., Cote, L. R., Haynes, O. M., Hahn, C. S., & Park, Y. (2010). Parenting knowledge: Experiential and sociodemographic factors in European American mothers of young children. Developmental Psychology, 46, 1677–1693. Bornstein, M. H., Cote, L. R., Haynes, O. M., Suwalsky, J. T., & Bakeman, R. (2012). Modalities of infant–mother interaction in Japanese, Japanese American immigrant, and European American dyads. Child Development, 83(6), 2073–2088. doi: 10.1111/j.1467– 8624.2012.01822.x Bornstein, M. H., Cote, L. R., Maital, S., Painter, K., Park, S. Y., Pascual, L.,… Vyt, A. (2004). Cross-linguistic analysis of vocabulary in young children: Spanish, Dutch, French, Hebrew, Italian, Korean, and American English. Child Development, 75(4), 1115–1139. Bornstein, M. H., Cote, L. R., & Venuti, P. (2001). Parenting beliefs and behaviors in northern

and southern groups of Italian mothers of young infants. Journal of Family Psychology, 15, 663–675. Bornstein, M. H., Gini, M., Putnick, D. L., Haynes, O. M., Painter, K. M., & Suwalsky, J. T. D. (2006). Short-term reliability and continuity of emotional availability in mother–child dyads across contexts of observation. Infancy, 10, 1–16. Bornstein, M. H., Hahn, C. S., & Haynes, O. M. (2011). Maternal personality, parenting cognitions, and parenting practices. Developmental Psychology, 47(3), 658–675. doi: 10.1037/a0023181 Bornstein, M. H., Hahn, C. S., Haynes, O. M., Belsky, J., Azuma, H., Kwak, K.,… de Galperín, C. Z. (2007). Maternal personality and parenting cognitions in cross-cultural perspective. International Journal of Behavioral Development, 31(3), 193–209. doi: 10.1177/0165025407074632 Bornstein, M. H., Hahn, C. S., Suwalsky, J. T. D., & Haynes, O. M. (2003). Socioeconomic status, parenting, and child development: The Hollingshead Four-Factor Index of social status and the socioeconomic index of occupations. In M. H. Bornstein & R. H. Bradley (Eds.), Socioeconomic Status, Parenting, and Child Development (pp. 29–82). Mahwah, NJ: Erlbaum. Bornstein, M. H., Haynes, O. M., Pascual, L., & Painter, K. M. (2004). Competence and satisfaction in parenting young children: An ecological, multivariate comparison of expressions and sources of self-evaluation in the United States and Argentina. In U. P. Gielen & J. L. Roopnarine (Eds.), Childhood and Adolescence: Cross-cultural Perspectives and Applications (pp. 166–195). Westport, CT: Praeger/Greenwood Press. Bornstein, M. H., Hendricks, C., Hahn, C. S., Haynes, O. M., Painter, K. M., & TamisLeMonda, C. S. (2003). Contributors to self-perceived competence, satisfaction, investment, and role balance in maternal parenting: A multivariate ecological analysis. Parenting: Science and Practice, 3(4), 285–326. Bornstein, M. H., Hendricks, C., Haynes, O. M., & Painter, K. M. (2007). Maternal sensitivity and child responsiveness: Associations with social context, maternal characteristics, and child characteristics in a multivariate analysis. Infancy, 12, 189–223. Bornstein, M. H., Jager, J., & Steinberg, L. D. (2012). Adolescents, parents, friends/peers: A relationships model (with commentary and illustrations). In M. A. Easterbrooks, J. Mistry, R. M. Lerner, & I. Weiner (Eds.), Handbook of psychology: Developmental psychology (2nd ed., Vol. 6, pp. 393–433). Hoboken, NJ: Wiley. Bornstein, M. H., & Lansford, J. E. (2010). Parenting. In M. H. Bornstein (Ed.), The handbook of cultural developmental science, Part 1: Domains of development across cultures (pp. 259–277). New York, NY: Psychology Press.

Bornstein, M. H., Park, Y., Putnick, D. L., Suwalsky, J. T. D., & Haynes, O. M. (2015). Infancy and parenting in 11 societies around the world—North, East, South, and West: Argentina, Belgium, Brazil, Cameroon, France, Israel, Italy, Japan, Kenya, South Korea, and the United States. Unpublished manuscript, Eunice Kennedy Shriver National Institute of Child Health and Human Development, Bethesda, MD. Bornstein, M. H., & Putnick, D. L. (2007). Chronological age, cognitions, and practices in European American mothers: A multivariate study of parenting. Developmental Psychology, 43(4), 850–864. doi: 10.1037/0012–1649.43.4.850 Bornstein, M. H., & Putnick, D. L. (2012). Cognitive and socioemotional caregiving in developing countries. Child Development, 83(1), 46–61. doi: 10.1111/j.1467– 8624.2011.01673.x Bornstein, M. H., & Putnick, D. L. (2015). IV. Mothering and fathering daughters and sons in low- and middle-income countries. In M. H. Bornstein, D. L. Putnick, J. E. Lansford, K. Deater-Deckard, & R. H. Bradley, Gender in low- and middle-income countries. Monograph of the Society for Research in Child Development. Bornstein, M. H., Putnick, D. L., Bradley, R. H., Lansford, J. E., & Deater-Deckard, K. (2015). Pathways among caregiver education, household resources, and infant growth in 39 low- and middle-income countries. Infancy, 20, 353–376. Bornstein, M. H., Putnick, D. L., Cote, L. R., Haynes, O. M., & Suwalsky, J. T. (2015). Mother-infant contingent vocalizations in 11 countries. Psychological Science, 26(8), 1272– 1284. doi: 10.1177/0956797615586796 Bornstein, M. H., Putnick, D. L., Heslington, M., Gini, M., Suwalsky, J. T. D., Venuti, P.,… Zingman de Galperín, C. (2008). Mother-child emotional availability in ecological perspective: Three countries, two regions, two genders. Developmental Psychology, 44, 666– 680. doi: 10.1037/0012–1649.44.3.666 Bornstein, M. H., Putnick, D. L., & Lansford, J. E. (2011). Parenting attributions and attitudes in cross-cultural perspective. Parenting: Science and Practice, 11(2–3), 214–237. doi: 10.1080/15295192.2011.585568 Bornstein, M. H., Putnick, D. L., Lansford, J. E., Deater-Deckard, K., & Bradley, R. (2015a). Gender in low- and middle-income countries. Monographs of the Society for Research in Child Development. Bornstein, M. H., Putnick, D. L., Lansford, J. E., Pastorelli, C., Skinner, A. T., Sorbring, E.,… Oburu, P. (2015b). Mother and father socially desirable responding in nine countries: Two kinds of agreement and relations to parenting self-reports. International Journal of Psychology, 50, 174–185. Bornstein, M. H., Putnick, D. L., & Suwalsky, J. T. D. (2012). A longitudinal process analysis

of mother–child emotional relationships in a rural Appalachian European American community. American Journal of Community Psychology, 50(1–2), 89–100. doi: 10.1007/s10464–011–9479–1 Bornstein, M. H., Putnick, D. L., Suwalsky, J. T. D., & Gini, M. (2006). Maternal chronological age, prenatal and perinatal history, social support, and parenting of infants. Child Development, 77(4), 875–892. doi: 10.1111/j.1467–8624.2006.00908.x Bornstein, M. H., & Ruddy, M. (1984). Infant attention and maternal stimulation: Prediction of cognitive and linguistic development in singletons and twins. In H. Bouma & D. Bouwhuis (Eds.), Attention and performance: Control of language processes (pp. 433–445). Hove, UK: Erlbaum. Bornstein, M. H., & Sawyer, J. (2005). Family systems. In K. McCartney & D. Phillips (Eds.), Blackwell handbook of early childhood development (pp. 381–398). Malden, MA: Blackwell. doi: 10.1002/9780470757703.ch19 Bornstein, M. H., & Suess, P. E. (2000) Child and mother cardiac vagal tone: Continuity, stability, and concordance across the first 5 years. Developmental Psychology, 36(1), 54–65. Bornstein, M. H., Tal, J., Rahn, C., Galperin, C. Z., Pecheux, M. G., Lamour, M.,… Tamis LeMonda, C. S. (1992). Functional analysis of the contents of maternal speech to infants of 5 and 13 months in four cultures: Argentina, France, Japan, and the United States. Developmental Psychology, 28(4), 593–603. doi: 10.1037/0012–1649.28.4.593 Bornstein, M. H., & Tamis-LeMonda, C. S. (1990). Activities and interactions of mothers and their firstborn infants in the first six months of life: Covariation, stability, continuity, correspondence, and prediction. Child Development, 61(4), 1206–1217. Bornstein, M. H., Tamis-LeMonda, C. S., Hahn, C. S., & Haynes, O. M. (2008). Maternal responsiveness to very young children at three ages: Longitudinal analysis of a multidimensional modular and specific parenting construct. Developmental Psychology, 44, 867–874. Bornstein, M. H., & Toole, M. (2010). Assessment of parenting. In S. Tyano, M. Keren, H. Herrman, & J. Cox (Eds.), Parenthood and Mental Health: A Bridge between Infant and Adult Psychiatry (pp. 349–355). Chichester, West Sussex, U.K.: Wiley-Blackwell. Borre, A., & Kliewer, W. (2014). Parental strain, mental health problems, and parenting practices: A longitudinal study. Personality and Individual Differences, 68, 93–97. Bosquet, M., & Egeland, B. (2000). Predicting parenting behaviors from antisocial practices content scale scores of the MMPI-2 administered during pregnancy. Journal of Personality Assessment, 74(1), 146 –162. doi: 10.1207/S15327752JPA740110 Boström, P. K., Broberg, M., & Hwang, P. (2010). Parents' descriptions and experiences of young children recently diagnosed with intellectual disability. Child: Care, Health and

Development, 36(1), 93–100. doi: 10.1111/j.1365–2214.2009.01036.x Boutelle, K. N., Cafri, G., & Crow, S. J. (2011). Parent-only treatment for childhood obesity: A randomized controlled trial. Obesity, 19(3), 574–580. doi: 10.1038/oby.2010.238 Bower-Russa, M. E., Knutson, J. F., & Winebarger, A. (2001). Disciplinary history, adult disciplinary attitudes, and risk for abusive parenting. Journal of Community Psychology, 29, 219–240. Bowlby, J. (1969). Attachment and loss: Attachment (2nd ed., Vol. 1). New York, NY: Basic Books. Boyd, R., & Richerson, P. J. (2005). The origin and evolution of cultures. New York, NY: Oxford University Press. Bradley, R. H., & Caldwell, B. M. (1995). The acts and conditions of the caregiving environment. Developmental Review, 15(1), 92–96. Bradley, R. H., & Corwyn, R. F. (2002). Socioeconomic status and child development. Annual Review of Psychology, 53(1), 371–399. doi: 10.1146/annurev.psych.53.100901.135233 Bradley, R. H., Corwyn, R. F., McAdoo, H. P., & García Coll, C. (2001). The home environments of children in the United States, Part I: Variations by age, ethnicity, and poverty status. Child Development, 72(6), 1844–1867. doi: 10.1111/1467–8624.t01–1–00382 Brady, E., & Guerin, S. (2010). “Not the romantic, all happy, coochy coo experience”: A qualitative analysis of interactions on an Irish parenting web site. Family Relations, 59(1), 14–27. doi: 10.1111/j.1741–3729.2009.00582.x Braungart-Rieker, J. M., Hill-Soderlund, A. L., & Karrass, J. (2010). Fear and anger reactivity trajectories from 4 to 16 months: The roles of temperament, regulation, and maternal sensitivity. Developmental Psychology, 46(4), 791–804. doi: 10.1037/a0019673 Bridges, R. S. (2008). Neurobiology of the parental brain. Waltham, MA: Academic Press. Bridgett, D. J., Laake, L. M., Gartstein, M. A., & Dorn, D. (2013). Development of infant positive emotionality: The contribution of maternal characteristics and effects on subsequent parenting. Infant and Child Development, 22(4), 362–382. doi: 10.1002/icd.1795 Briga, M., Pen, I., & Wright, J. (2012). Care for kin: Within-group relatedness and allomaternal care are positively correlated and conserved throughout the mammalian phylogeny. Biology Letters, 8(4), 533–536. doi: 10.1098/rsbl.2012.0159 Briggs, X. S., Popkin, S. J., & Goering, J. (2010). Moving to opportunity: The story of an American experiment to fight ghetto poverty. New York, NY: Oxford University Press. Brody, G. H., & Flor, D. L. (1998). Maternal resources, parenting practices, and child competence in rural, single-parent African American families. Child Development, 69(3),

803–816. Brody, G. H., Ge, X., Kim, S. Y., Murry, V. M., Simons, R. L., Gibbons, F. X.,… Conger, R. D. (2003). Neighborhood disadvantage moderates associations of parenting and older sibling problem attitudes and behavior with conduct disorders in African American children. Journal of Consulting and Clinical Psychology, 71(2), 211. doi: 10.1037/0022–006X.71.2.211 Bronfenbrenner, U. (1958). Socialization and social class through time and space. New York, NY: Holt. Bronfenbrenner, U. (1979). Contexts of child rearing: Problems and prospects. American Psychologist, 34(10), 844. Bronfenbrenner, U., & Crouter, A. C. (1983). The evolution of environmental models in developmental research. In P. H. Mussen (Series Ed.) & W. Kessen (Vol. Ed.), Handbook of child psychology: History, theory, and methods (4th ed., Vol. 1, pp. 357–414). New York, NY: Wiley. Bronfenbrenner, U., & Morris, P. A. (2006). The bioecological model of human development. In W. Damon & R. M. Lerner (Series Eds.) & R. L. Lerner (Vol. Ed.), Handbook of child psychology: Theoretical models of human development (6th ed., Vol. 1, pp. 793–828). Hoboken, NJ: Wiley. doi: 10.1002/9780470147658.chpsy0114 Brook, J. S., Brook, D. W., Ning, Y., Whiteman, M., & Finch, S. J. (2006). The relationship of personality and behavioral development from adolescence to young adulthood and subsequent parenting behavior. Psychological Reports, 99(1), 3–19. Brown, J., Cohen, P., Johnson, J. G., & Salzinger, S. (1998). A longitudinal analysis of risk factors for child maltreatment: Findings of a 17-year prospective study of officially recorded and self-reported child abuse and neglect. Child Abuse & Neglect, 22(11), 1065–1078. Brown, L. D., Goslin, M. C., & Feinberg, M. E. (2011). Relating engagement to outcomes in prevention: The case of a parenting program for couples. American Journal of Community Psychology, 50(1–2), 17–25. doi: 10.1007/s10464–011–9467–5 Browne, D. T., Meunier, J. C., O'Connor, T. G., & Jenkins, J. M. (2012). The role of parental personality traits in differential parenting. Journal of Family Psychology, 26, 542–553. Browning, C. R., Leventhal, T., & Brooks-Gunn, J. (2005). Sexual initiation in early adolescence: The nexus of parental and community control. American Sociological Review, 70(5), 758–778. doi: 10.1177/000312240507000502 Brunton, P. J., & Russell, J. A. (2008). The expectant brain: Adapting for motherhood. Nature Reviews Neuroscience, 9(1), 11–25. doi: 10.1038/nrn2280 Buehler, C., & O'Brien, M. (2011). Mothers' part-time employment: Associations with mother and family well-being. Journal of Family Psychology, 25(6), 895. doi: 10.1037/a0025993

Brueckner, M., & Schwandt, H. (2014). Income and population growth. Economics Journal (IZA Discussion Paper No. 7422), 1–41. doi: 10.1111/ecoj.12152 Bugental, D. B., & Happaney, K. (2002). Parental attributions. In M. H. Bornstein (Ed.), Handbook of parenting: Being and becoming a parent (2nd ed., Vol. 3, pp. 509–535). Mahwah, NJ: Erlbaum. Bugental, D. B., & Happaney, K. (2004). Predicting infant maltreatement in low-income families: The interactive effects of maternal attributions and child status at birth. Developmental Psychology, 40, 234–243. doi: 10.1037/0012–1649.40.2.234 Buhrmester, D., Camparo, L., Christensen, A., Gonzalez, L. S., & Hinshaw, S. P. (1992). Mothers and fathers interacting in dyads and triads with normal and hyperactive sons. Developmental Psychology, 28, 500–509. Bumpus, M. F., Crouter, A. C., & McHale, S. M. (2006). Linkages between negative work-tofamily spillover and mothers' and fathers' knowledge of their young adolescents' daily lives. Journal of Early Adolescence, 26, 36–59. Burch, R. L., & Gallup, G. G., Jr. (2000). Perceptions of paternal resemblance predict family violence. Evolution and Human Behavior, 21(6), 429–435. doi: 10.1016/S1090– 5138(00)00056–8 Burchi, F. (2012). Whose education affects a child's nutritional status? From parents' to household's education. Demographic Research, 27(23), 681–704. doi: 10.4054/DemRes.2012.27.23 Burchinal, M., Magnuson, K., Powell, D., & Hong, S. S. (2015). Early child care and education. In M. H. Bornstein & T. Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R. M. Lerner (Ed.), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 223–267). Hoboken, NJ: Wiley. Burchinal, M., Skinner, D., & Reznick, S. (2010). European American and African American mothers' beliefs about parenting and disciplining infants: A mixed-method analysis. Parenting: Science and Practice, 10(2), 79–96. doi: 10.1080/15295190903212604 Burchinal, M., Vernon-Feagans, L., Cox, M., & Key Family Life Project Investigators. (2008). Cumulative social risk, parenting, and infant development in rural low-income communities. Parenting: Science and Practice, 8, 41–69. doi: 10.1080/15295190701830672 Burkhouse, K. L., Gibb, B. E., Coles, M. E., Knopik, V. S., & McGeary, J. E. (2011). Serotonin transporter genotype moderates the link between children's reports of overprotective parenting and their behavioral inhibition. Journal of Abnormal Child Psychology, 39(6), 783–790. doi: 10.1007/s10802–011–9526–2 Burton, L., & Jarrett, R. L. (2000). In the mix yet on the margins: The place of families in urban neighborhood and child development research. Journal of Marriage and Family, 62(4), 1111–

1135. doi: 10.1111/j.1741–3737.2000.01114.x Butterfield, E. C., Albertson, L. R., & Johnston, J. C. (1995). On making cognitive theory more general and developmentally pertinent. In F. E. Weinert & W. Schneider (Eds.), Memory performance and competencies: Issues in growth and development (pp. 181–205). Hillsdale, NJ: Erlbaum. Calzada, E. J., Eyberg, S. M., Rich, B., & Querido, J. G. (2004). Parenting disruptive preschoolers: Experiences of mothers and fathers. Journal of Abnormal Child Psychology, 32(2), 203–213. doi: 10.1023/B:JACP.0000019771.43161.1c Campbell, J., & Gilmore, L. (2007). Intergenerational continuities and discontinuities in parenting styles. Australian Journal of Psychology, 59(3), 140–150. doi: 10.1080/00049530701449471 Campbell, S. B., Matestic, P., von Stauffenberg, C., Mohan, R., & Kirchner, T. (2007). Trajectories of maternal depressive symptoms, maternal sensitivity, and children's functioning at school entry. Developmental Psychology, 43(5), 1202. doi: 10.1037/0012–1649.43.5.1202 Campos, J. J., Anderson, D. I., Barbu-Roth, M. A., Hubbard, E. M., Hertenstein, M. J., & Witherington, D. (2000). Travel broadens the mind. Infancy, 1(2), 149–219. doi: 10.1207/S15327078IN0102_1 Cannon, E. A., Schoppe-Sullivan, S. J., Mangelsdorf, S. C., Brown, G. L., & Sokolowski, M. S. (2008). Parent characteristics as antecedents of maternal gatekeeping and fathering behavior. Family Process, 47(4), 501–519. doi: 10.1111/j.1545–5300.2008.00268.x Capaldi, D. M., Pears, K. C., Patterson, G. R., & Owen, L. D. (2003). Continuity of parenting practices across generations in an at-risk sample: A prospective comparison of direct and mediated associations. Journal of Abnormal Child Psychology, 31(2), 127–142. doi: 10.1023/A:1022518123387 Carew, J. V. (1980). Experience and the development of intelligence in young children at home and in day care. Monographs of the Society for Research in Child Development, 45, 6–7 (Serial No. 187). Caria, A., de Falco, S., Venuti, P., Lee, S., Esposito, G., Rigo, P.,… Bornstein, M. H. (2012). Species-specific response to human infant faces in the premotor cortex. NeuroImage, 60(2), 884–893. doi: 10.1016/j.neuroimage.2011.12.068 Carpiano, R. M., & Kimbro, R. T. (2012). Neighborhood social capital, parenting strain, and personal mastery among female primary caregivers of children. Journal of Health and Social Behavior, 53(2), 232–247. doi: 10.1177/0022146512445899 Casey, R. J., & Ritter, J. M. (1996). How infant appearance informs: Child care providers' responses to babies varying in appearance of age and attractiveness. Journal of Applied Developmental Psychology, 17(4), 495–518. doi: 10.1016/S0193–3973(96)90013–1

Caspi, A., & Elder, G. H., Jr. (1988). Emergent family patterns: The intergenerational construction of problem behaviour and relationships. In R. A. Hinde & J. Stevenson-Hinde (Eds.), Relationships within families: Mutual influences (pp. 218–240). Oxford, UK: Clarendon Press. Caspi, A., McClay, J., Moffitt, T. E., Mill, J., Martin, J., Craig, I. W.,… Poulton, R. (2002). Role of genotype in the cycle of violence in maltreated children. Science, 297(5582), 851– 854. doi: 10.1126/science.1072290 Caspi, A., Moffitt, T. E., Morgan, J., Rutter, M., Taylor, A., Arseneault, L.,… Polo-Tomas, M. (2004). Maternal expressed emotion predicts children's antisocial behavior problems: Using monozygotic-twin differences to identify environmental effects on behavioral development. Developmental Psychology, 40(2), 149–161. doi: 10.1037/0012–1649.40.2.149 Caspi, A., Sugden, K., Moffitt, T. E., Taylor, A., Craig, I. W., Harrington, H.,… Poulton, R. (2003). Influence of life stress on depression: Moderation by a polymorphism in the 5-HTT gene. Science Signaling, 301(5631), 386. doi: 10.1126/science.108396 Caspi, A., Taylor, A., Moffitt, T. E., & Plomin, R. (2000). Neighborhood deprivation affects children's mental health. Psychological Science, 11(4), 338–342. doi: 10.1111/1467– 9280.00267 Catron, T. R., & Masters, J. C. (1993). Mothers' and children's conceptualizations of corporal punishment. Child Development, 64(6), 1815–1828. Caudill, W., & Weinstein, H. (1969). Maternal care and infant behavior in Japan and America. Psychiatry, 32, 12–43. Caumont, A., & Wang, W. (2014). 5 questions (and answers) about American moms today. Retrieved from Pew research center website: http://www.pewresearch.org/facttank/2014/05/09/5-questions-and-answers-about-american-moms-today/ Cavanagh, K., Dobash, R. E., & Dobash, R. P. (2007). The murder of children by fathers in the context of child abuse. Child Abuse and Neglect, 31(7), 731–746. doi: 10.1016/j.chiabu.2006.12.016 CDC/National Center for Health Statistics. (2015). http://www.cdc.gov/nchs/fastats/births.htm Champagne, F. A., Francis, D. D., Mar, A., & Meaney, M. J. (2003). Variations in maternal care in the rat as a mediating influence for the effects of environment on development. Physiology & Behavior, 79(3), 359–371. Champagne, F. A. (2008). Epigenetic mechanisms and the transgenerational effects of maternal care. Frontiers in Neuroendocrinology, 29(3), 386–397. doi: 10.1016/j.yfrne.2008.03.003 Champagne, F. A. (2010). Epigenetic influence of social experiences across the lifespan. Developmental Psychobiology, 52(4), 299–311. doi: 10.1002/dev.20436

Champagne, F., Diorio, J., Sharma, S., & Meaney, M. J. (2001). Naturally occurring variations in maternal behavior in the rat are associated with differences in estrogen-inducible central oxytocin receptors. Proceedings of the National Academy of Sciences USA, 98(22), 12736– 12741. doi: 10.1073/pnas.221224598 Champagne, F. A., & Meaney, M. J. (2007). Transgenerational effects of social environment on variations in maternal care and behavioral response to novelty. Behavioral Neuroscience, 121(6), 1353. doi: 10.1037/0735–7044.121.6.1353 Champion, J. E., Jaser, S. S., Reeslund, K. L., Simmons, L., Potts, J. E., Shears, A. R., & Compas, B. E. (2009). Caretaking behaviors by adolescent children of mothers with and without a history of depression. Journal of Family Psychology, 23(2), 156–166. doi: 10.1037/a0014978 Chao, R., & Tseng, V. (2002). Parenting of Asians. In M. H. Bornstein (Ed.), Handbook of parenting: Social conditions and applied parenting (2nd ed., Vol. 4, pp. 59–93). Mahwah, NJ: Erlbaum. Charney, E. (2012). Behavior genetics and postgenomics. Behavioral and Brain Sciences, 35(5), 331–358. doi: 10.1017/S0140525X11002226 Chen, X., Cen, G., Li, D., & He, Y. (2005). Social functioning and adjustment in Chinese children: The imprint of historical time. Child Development, 76(1), 182–195. doi: 10.1111/j.1467–8624.2005.00838.x Chen, X., Hastings, P. D., Rubin, K. H., Chen, H., Cen, G., & Stewart, S. L. (1998). Child rearing attitudes and behavioral inhibition in Chinese and Canadian toddlers: A cross-cultural study. Developmental Psychology, 34, 677–686. doi: 10.1037/0012–1649.34.4.677 Chen, Z., & Kaplan, H. B. (2001). Intergenerational transmission of constructive parenting. Journal of Marriage and Family, 63(1), 17–31. doi: 10.1111/j.1741–3737.2001.00017.x Cherney, I. D. (2010). Mothers', fathers' and their children's perceptions and reasoning about nurturance and self-determination rights. International Journal of Children's Rights, 18(1), 79–99. doi: 10.1163/092755609X12482670820520 Chilcoat, H. D., & Anthony, J. C. (1996). Impact of parent monitoring on initiation of drug use through late childhood. Journal of the American Academy of Child & Adolescent Psychiatry, 35(1), 91–100. Child Trends. (2013). Attitudes toward spanking. Retrieved from http://www.childtrends.org/? indicators=attitudes-toward-spanking Chu, J. T., Farruggia, S., Sanders, M. R., & Ralph, A. (2012). Towards a public health approach to parenting programmes for parents of adolescents. Journal of Public Health, 34, 41–47. doi: 10.1093/pubmed/fdr123

Chuang, Y. C., Ennett, S. T., Bauman, K. E., & Foshee, V. A. (2005). Neighborhood influences on adolescent cigarette and alcohol use: Mediating effects through parent and peer behaviors. Journal of Health and Social Behavior, 46(2), 187–204. doi: 10.1177/002214650504600205 Cicchetti, D. (2010). Resilience under conditions of extreme stress: A multilevel perspective. World Psychiatry, 9(3), 145–154. doi: 10.1002/j.2051–5545.2010.tb00297.x Cicchetti, D., & Toth, S. L. (2009). The past achievements and future promises of developmental psychopathology: The coming of age of a discipline. Journal of Child Psychology and Psychiatry, 50 (1–2), 16–25. Cicchetti, D., Toth, S. L., & Maughan, A. (2000). An ecological-transactional model of child maltreatment. In A. J. Sameroff, M. Lewis, & S. M. Miller (Eds.), Handbook of developmental psychopathology (2nd ed., pp. 689–722). New York, NY: Kluwer Academic Press/Plenum Press. Ciciolla, L., Crnic, K. A., & West, S. G. (2012). Determinants of change in maternal sensitivity: Contributions of context, temperament, and developmental risk. Parenting: Science and Practice, 13(3), 178–195. doi: 10.1080/15295192.2013.756354 Civitas Initiative, Zero to Three, & Brio Corporation. (2000). What grown-ups understand about child development: A national benchmark survey. Washington, DC: Zero to Three, National Center for Infants, Toddlers, and Families. Clark, L. A., Kochanska, G., & Ready, R. (2000). Mothers' personality and its interaction with child temperament as predictors of parenting behavior. Journal of Personality and Social Psychology, 79(2), 274–285. doi: 10.1037/0022–3514.79.2.274 Clarke-Stewart, K. A. (1973). Interactions between mothers and their young children: Characteristics and consequences. Monographs of the Society for Research in Child Development, 1–109. Clarke-Stewart, K. A. (1998). Historical shifts and underlying themes in ideas about rearing young children in the United States: Where have we been? Where are we going? Early Development and Parenting, 7, 101–117. Clarke-Stewart, K. A., & Allhusen, V. D. (2002). Nonparental caregiving. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 215–252). Mahwah, NJ: Erlbaum. Clément, M.-È., & Chamberland, C. (2009). The role of parental stress, mother's childhood abuse and perceived consequences of violence in predicting attitudes and attribution in favor of corporal punishment. Journal of Child and Family Studies, 18, 163–171. doi: 10.1007/s10826-008-9216-z Clifford, E. (1959). Discipline in the home: A controlled observational study of parental practices. Journal of Genetic Psychology, 95, 45–82.

Clutton-Brock, T. H. (1989). Review lecture: Mammalian mating systems. Proceedings of the Royal Society of London. B. Biological Sciences, 236(1285), 339–372. Clutton-Brock, T. H. (1991). The evolution of parental care. Princeton, NJ: Princeton University Press. Clutton-Brock, T. H., & Vincent, A. C. (1991). Sexual selection and the potential reproductive rates of males and females. Nature, 351(6321), 58–60. Cochran, M. (1993). Parenting and personal social networks. In T. Luster & L. Okagaki (Eds.), Parenting: An ecological perspective (pp. 149–178). Hillsdale, NJ: Erlbaum. Cochran, M., & Niego, S. (2002). Parenting and social networks. In M. H. Bornstein (Ed.), Handbook of parenting: Applied parenting (2nd ed., Vol. 4, pp. 123–148). Mahwah, NJ: Erlbaum. Cohler, B. J., & Paul, S. (2002). Psychoanalysis and parenthood. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 563– 599). Mahwah, NJ: Erlbaum. Cohn, D., & Livingston, G. (2013). Record share of new mothers are college educated. Retrieved from Pew Research Center website: http://www.pewsocialtrends.org/files/2013/05/fertilityeducation_final.pdf Cole, M. (1996). Cultural psychology: A once and future discipline. Cambridge, MA: Harvard University Press. Coleman, P. K., & Karraker, K. H. (1998). Self-efficacy and parenting quality: Findings and future applications. Developmental Review, 18, 47–85. Coleman, P. K., & Karraker, K. H. (2003). Maternal self-efficacy beliefs, competence in parenting, and toddlers' behavior and developmental status. Infant Mental Health Journal, 24(2), 126–148. doi: 10.1002/imhj.10048 Coley, R. L., & Chase-Lansdale, P. L. (1998). Adolescent pregnancy and parenthood: Recent evidence and future directions. American Psychologist, 53, 152–166. Collett, B., Gimpel, G., Greenson, J., & Gunderson, T. (2001). Assessment of discipline styles among parents of preschool through school-age children. Journal of Psychopathology and Behavioral Assessment, 23(3), 163–170. doi: 10.1023/A:1010965220517 Collins, W. A., Maccoby, E. E., Steinberg, L., Hetherington, E. M., & Bornstein, M. H. (2000). Contemporary research on parenting: The case for nature and nurture. American Psychologist, 55(2), 218–232. doi: 10.1037/0003–066X.55.2.218 Collins, W. A., & Russell, G. (1991). Mother-child and father-child relationships in middle childhood and adolescence: A developmental analysis. Developmental Review, 11, 99–136.

Condry, J., & Condry, S. (1976). Sex differences: A study of the eye of the beholder. Child Development, 47(3), 812–819. Conger, R. D. (2013). Rural children at risk. Monographs of the Society for Research in Child Development, 78, 127–138. doi: 10.1111/mono.12055 Conger, R. D., Belksy, J., & Capaldi, D. M. (2009). The intergenerational transmission of parenting: Closing comments for the special section. Developmental Psychology, 45, 1276– 1283. doi: 10.1037/a0016911 Conger, R. D., & Donnellan, M. B. (2007). An interactionist perspective on the socioeconomic context of human development. Annual Review of Psychology, 58, 175–199. doi: 10.1146/annurev.psych.58.110405.085551 Conger, R. D., & Ge, X. (1999). Conflict and cohesion in parent–adolescent relations: Changes in emotional expression from early to midadolescence. In M. J. Cox & J. Brooks-Gunn (Eds.), Conflict and cohesion in families (pp. 185–206). Mahwah, NJ: Erlbaum. Conger, R. D., Ge, X., Elder, G. H., Jr., Lorenz, F. O., & Simons, R. L. (1994). Economic stress, coercive family process, and developmental problems of adolescents. Child Development, 65, 541–561. Conger, R. D., Schofield, T. J., & Neppl, T. K. (2012). Intergenerational continuity and discontinuity in harsh parenting. Parenting: Science and Practice, 12(2–3), 222–231. doi: 10.1080/15295192.2012.683360 Conger, R. D., & Simons, R. L. (1997). Life-course contingencies in the development of adolescent antisocial behavior: A matching law approach. Advances in Criminological Theory, 7, 55–99. Conrad, B., Gross, D., Fogg, L., & Ruchala, P. (1992). Maternal confidence, knowledge, and quality of mother-toddler interactions: A preliminary study. Infant Mental Health Journal, 13, 353–362. Coplan, R. J., Hastings, P. D., Lagacé-Séguin, D. G., & Moulton, C. E. (2002). Authoritative and authoritarian mother's parenting goals, attributions, and emotions across different childrearing contexts. Parenting: Science and Practice, 2(1), 1–26. doi: 10.1207/S15327922PAR0201_1 Corpaci, F., & Wachs, T. D. (2002). Does parental mood or efficacy mediate the influence of environmental chaos upon parenting behavior? Merrill-Palmer Quarterly, 48, 182–201. doi: 10.1353/mpq.2002.0006 Corter, C. M., & Fleming, A. S. (2002). Psychobiology of maternal behavior in human beings. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 141–181). Mahwah, NJ: Erlbaum.

Costigan, C. L., Cox, M. J., & Cauce, A. M. (2003). Work-parenting linkages among dualearner couples at the transition to parenthood. Journal of Family Psychology, 17, 397–408. Cote, L. R., & Bornstein, M. H. (2000). Social and didactic parenting behaviors and beliefs among Japanese American and South American mothers of infants. Infancy, 1(3), 363–374. doi: 10.1207/S15327078IN0103_5 Cote, L. R., & Bornstein, M. H. (2003). Cultural and parenting cognitions in acculturating cultures: I. Cultural comparisons and developmental continuity and stability. Journal of Cross-Cultural Psychology, 34, 323–349. Cote, L. R., & Bornstein, M. H. (2005). Japanese American and South American immigrant mothers' perceptions of their own and their spouses' parenting styles. In H. Grietens, W. Lahaye, W. Hellinckx, & L. Vandemeulebroecke (Eds.), In The Best Interests of Children and Youth: International Perspectives (pp. 47–76). Leuven, Belgium: Leuven University Press (Series Studia Paedagogica). Cote, L. R., Bornstein, M. H., Haynes, O. M., & Bakeman, R. (2008). Mother-infant personand object-directed interactions in Latino immigrant families: A comparative approach. Infancy, 13, 338–365. doi: 10.1080/15250000802189386 Cote, L. R., Kwak, K., Putnick, D. L., Chung, H. J., & Bornstein, M. H. (2015). The acculturation of parenting cognitions: A comparison of South Korean, Korean Immigrant, and European American mothers. Journal of Cross-Cultural Psychology. Cowan, C. P., & Cowan, P. A. (1992). When partners become parents. New York, NY: Basic Books. Cowan, P. A., & Cowan, C. P. (2002). Interventions as tests of family systems theories: Marital and family relationships in children's development and psychopathology. Development and Psychopathology, 14(4), 731–759. doi: 10.1017/S0954579402004054 Cox, M. J., & Harter, K. S. M. (2003). Parent-child relationships. In M. H. Bornstein, L. Davidson, C. L. M. Keyes, & K. A. Moore (Eds.), Well-being: Positive development across the life course (pp. 191–204). Mahwah, NJ: Erlbaum. Cox, M. J., & Paley, B. (2003). Understanding families as systems. Current Directions in Psychological Science, 12(5), 193–196. doi: 10.1111/1467–8721.01259 Coyl, D. D., Roggman, L. A., & Newland, L. A. (2002). Stress, maternal depression, and negative mother-infant interactions in relation to infant attachment. Infant Mental Health Journal, 23(1–2), 145–163. doi: 10.1002/imhj.10009 Crittenden, A. (Ed.). (2004). If you've raised kids, you can manage anything: Leadership begins at home. New York, NY: Gotham Books. Crnic, K. A., Gaze, C., & Hoffman, C. (2005). Cumulative parenting stress across the

preschool period: Relations to maternal parenting and child behaviour at age 5. Infant and Child Development, 14(2), 117–132. doi: 10.1002/icd.384 Crnic, K. A., & Greenberg, M. T. (1990). Minor parenting stresses with young children. Child Development, 61(5), 1628–1637. Crnic, K. A., Greenberg, M. T., Ragozin, A. S., Robinson, N. M., & Basham, R. (1983). Effects of stress and social support on mothers and premature and full-term infants. Child Development, 54, 209–217. Crnic, K. A., & Low, C. (2002). Everyday stresses and parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Practical parenting (2nd ed., Vol. 5, pp. 243–267). Mahwah, NJ: Erlbaum. Crockenberg, S. B., (1986). Are temperamental differences in babies associated with predictable differences in care giving? New Directions for Child Development, 31, 53–73. Crockenberg, S. B. (1988). Social support and parenting. In W. Fitzgerald, B. Lester, & M. Yogman (Eds.), Research on support for parents and infants in the postnatal period (pp. 67– 92). New York, NY: Ablex. Crouter, A. C., Bumpus, M. F., Head, M. R., & McHale, S. M. (2001). Implications of overwork and overload for the quality of men's family relationships. Journal of Marriage and the Family, 63, 404–416. Crouter, A. C., & McHale, S. M. (1993). Temporal rhythms in family life: Seasonal variation in the relation between parental work and family processes. Developmental Psychology, 29, 198–205. Cummings, E. M., & Cummings, J. S. (2002). Parenting and attachment. In M. H. Bornstein (Ed.), Handbook of parenting: Practical parenting (2nd ed., Vol. 5, pp. 35–58). Mahwah, NJ: Erlbaum. Cunningham, F., Leveno, K., Bloom, S., Hauth, J., Rouse, D., & Spong, C. (2010). Williams obstetrics (23rd ed.). New York, NY: McGraw-Hill. Currie, J., & Schwandt, H. (2014). Short- and long-term effects of unemployment on fertility. Proceedings of the National Academy of Sciences, 111(41), 14734–14739. Cutrona, C. E., & Troutman, B. R. (1986). Social support, infant temperament, and parenting self-efficacy: A mediational model of postpartum depression. Child Development, 57, 1507– 1518. D'Onofrio, B. M., Rickert M. E., Frans E., Kuja-Halkola R., Almqvist C., Sjölander A.,… Lichtenstein P. (2014). Paternal age at childbearing and offspring psychiatric and academic morbidity. JAMA Psychiatry, 71(4), 432–438. doi: 10.1001/jamapsychiatry.2013.4525 Daddis, C., & Smetana, J. (2005). Middle-class African American families' expectations for

adolescents' behavioural autonomy. International Journal of Behavioral Development, 29(5), 371–381. doi: 10.1177/01650250500167053 Dadds, M. R., Maujean, A., & Fraser, J. A. (2003). Parenting and conduct problems in children: Australian data and psychometric properties of the Alabama Parenting Questionnaire. Australian Psychologist, 38(3), 238–241. doi: 10.1080/00050060310001707267 Dallaire, D. H., & Weinraub, M. (2005). The stability of parenting behaviors over the first 6 years of life. Early Childhood Research Quarterly, 20(2), 201–219. doi: 10.1016/j.ecresq.2005.04.008 Daly, M., & Wilson, M. (1988). Homicide. New York, NY: Aldine de Gruyter. Damashek, A., Doughty, D., Ware, L., Silovsky, J. (2011). Predictors of client engagement and attrition in home-based child maltreatment prevention services. Child Maltreatment, 16(1), 9– 20. doi: 10.1177/1077559510388507 Damato, E. G. (2005). Parenting multiple infants. Newborn and Infant Nursing Reviews, 5(4), 208–214. doi: 10.1053/j.nainr.2005.08.003 Dasgupta, S. D. (1998). Gender roles and cultural continuity in the Asian Indian immigrant community in the US. Sex Roles, 38(11–12), 953–974. Davis, E. F., Schoppe-Sullivan, S. J., Mangelsdorf, S. C., & Brown, G. L. (2009). The role of infant temperament in stability and change in coparenting across the first year of life. Parenting: Science and Practice, 9, 143–159. doi: 10.1080/15295190802656836 Davis, O., Haworth, C., Lewis, C., & Plomin, R. (2012). Visual analysis of geocoded twin data puts nature and nurture on the map. Molecular Psychiatry, 17, 867–874. doi: 10.1038/mp.2012.68 Davis, R. N., Davis, M. M., Freed, G. L., & Clark, S. J. (2011). Fathers' depression related to positive and negative parenting behaviours with 1-year old children. Pediatrics, 127, 612– 618. Dawkins, R. (1976). The selfish gene. Oxford, UK: Oxford University Press. Dayton, C. J., Levendosky, A. A., Davidson, W. S., & Bogat, G. A. (2010). The child as held in the mind of the mother: The influence of prenatal maternal representations on parenting behaviors. Infant Mental Health Journal, 31(2), 220–241. doi: 10.1002/imhj.20253 De Baca, T. C., Figueredo, A. J., & Ellis, B. J. (2012). An evolutionary analysis of variation in parental effort: Determinants and assessment. Parenting: Science and Practice, 12(2–3), 94– 104. doi: 10.1080/15295192.2012.680396 De Haan, A. D., Prinzie, P., & Deković, M. (2009). Mothers' and fathers' personality and parenting: The mediating role of sense of competence. Developmental Psychology, 45(6), 1695–1707. doi: 10.1037/a0016121

De Los Reyes, A., & Kazdin, A. E. (2005). Informant discrepancies in the assessment of childhood psychopathology: a critical review, theoretical framework, and recommendations for further study. Psychological Bulletin, 131(4), 483. doi: 10.1037/0033–2909.131.4.483 de Pisapia, N., Bornstein, M. H., Rigo, P., Esposito, G., de Falco, S., & Venuti, P. (2013). Sex differences in directional brain responses to infant hunger cries. NeuroReport, 24(3), 142– 146. doi: 10.1097/WNR.0b013e32835df4fa Deal, J. E., Hagan, M. S., Bass, B., Hetherington, E. M., & Clingempeel, G. (1999). Marital interaction in dyadic and triadic contexts: Continuities and discontinuities. Family Process, 38, 105–115. Dearing, E. (2004). The developmental implications of restrictive and supportive parenting across neighborhoods and ethnicities: Exceptions are the rule. Journal of Applied Developmental Psychology, 25(5), 555–575. doi: 10.1016/j.appdev.2004.08.007 Deater-Deckard, K. (1998). Parenting stress and child adjustment: Some old hypotheses and new questions. Clinical Psychology: Science and Practice, 5(3), 314–332. Deater-Deckard, K. (2004). Parenting stress. New Haven, CT: Yale University Press. Deater-Deckard, K., & Dodge, K. A. (1997). Externalizing behavior problems and discipline revisited: Nonlinear effects and variation by culture, context, and gender. Psychological Inquiry, 8, 161–175. DeBaryshe, B. D. (1995). Maternal belief systems: Linchpin in the home reading process. Journal of Applied Developmental Psychology, 16, 1–20. Decuyper, M., De Pauw, S., De Fruyt, F., De Bolle, M., & De Clercq, B. J. (2009). A meta analysis of psychopathy-, antisocial PD- and FFM associations. European Journal of Personality, 23(7), 531–565. DeGarmo, D. S., Eddy, J. M., Reid, J. B., & Fetrow, R. A. (2009). Evaluating mediators of the impact of the Linking the Interests of Families and Teachers (LIFT) multimodal preventive intervention on substance use initiation and growth across adolescence. Prevention Science, 10(3), 208–220. doi: 10.1007/s11121–009–0126–0 DeGarmo, D. S., & Forgatch, M. S. (2005). Early development of delinquency within divorced families: Evaluating a randomized preventive intervention trial. Developmental Science, 8, 229–239. doi: 10.1111/j.1467–7687.2005.00412.x Dekovic, M., & Meeus, W. (1997). Peer relations in adolescence: Effects of parenting and adolescent's self-concept. Journal of Adolescence, 20, 163–176. Del Vecchio, T., Walter, A., & O'Leary, S. G. (2009). Affective and physiological factors predicting maternal response to infant crying. Infant Behavior and Development, 32(1), 117– 122. doi: 10.1016/j.infbeh.2008.10.005

DeLongis, A., Capreol, M., Holtzman, S., O'Brien, T., & Campbell, J. (2004). Social support and social strain among husbands and wives: A multilevel analysis. Journal of Family Psychology, 18, 470–479. doi: 10.1037/0893–3200.18.3.470 DeMaris, A., Mahoney, A., & Pargament, K. (2011). Doing the scut work of infant care: Does religiousness encourage father involvement? Journal of Marriage and Family, 73(2), 354– 368. doi: 10.1111/j.1741–3737.2010.00811.x deMause, L. (Ed.). (1975). The new psychohistory. New York, NY: Psychohistory Press. Demers, I., Bernier, A., Tarabulsy, G. M., & Provost, M. A. (2010). Maternal and child characteristics as antecedents of maternal mind-mindedness. Infant Mental Health Journal, 31, 94–112. doi: 10.1002/imhj.20244 Demo, D. H., Small, S. A., & Savin-Williams, R. C. (1987). Family relations and the selfesteem of adolescents and their parents. Journal of Marriage and the Family, 705–715. Denckla, M. B., & Reiss, A. L. (1997). Prefrontal-subcortical circuits in developmental disorders. In N. A. Krasnegor, G. R. Lyon, & P. S. Goldman-Rakic (Eds.), Development of the prefrontal cortex: Evolution, neurobiology, and behavior (pp. 283–294). Baltimore, MD: Paul H. Brookes. Deutsch, F. M., Ruble, D. N., Fleming, A., Brooks-Gunn, J., & Stangor, C. (1988). Information seeking and maternal self-definition during the transition to motherhood. Journal of Personality and Social Psychology, 55, 420–431. Dias, B. G., & Ressler, K. J. (2014). Parental olfactory experience influences behavior and neural structure in subsequent generations. Nature Neuroscience, 17, 89–96. doi: 10.1038/nn.3594 Diel, M., Owen, S. K. & Youngblade, L. M. (2004). Agency and communion attributes in adults' spontaneous self-representations. International Journal of Behavioral Development, 28, 1–15. DiLillo, D., & Damashek, A. (2003). Parenting characteristics of women reporting a history of childhood sexual abuse. Child Maltreatment, 8(4), 319–333. doi: 10.1177/1077559503257104 Dishion, T. J., Nelson, S. E., & Bullock, B. M. (2004). Premature adolescent autonomy: Parent disengagement and deviant peer process in the amplification of problem behaviour. Journal of Adolescence, 27(5), 515–530. doi: 10.1016/j.adolescence.2004.06.005 Dishion, T. J., Poulin, F., & Medici Skaggs, N. (2000). The ecology of premature autonomy in adolescence: Biological and social influences. In K. A. Kerns, J. M. Contreras, & A. M. NealBarnett (Eds.), Family and peers: Linking two social worlds (pp. 27–45). Westport, CT: Praeger.

Dishion, T. J., Shaw, D., Connell, A., Gardner, F., Weaver, C., & Wilson, M. (2008). The Family Check-Up with high-risk indigent families: Preventing problem behavior by increasing parents' positive behavior support in early childhood. Child Development, 79(5), 1395–1414. doi: 10.1111/j.1467–8624.2008.01195.x Dix, T. (1991). The affective organization of parenting: Adaptive and maladaptive processes. Psychological Bulletin, 110, 3–25. Dix, T., & Meunier, L. N. (2009). Depressive symptoms and parenting competence: An analysis of 13 regulatory processes. Developmental Review, 29(1), 45–68. doi: 10.1016/j.dr.2008.11.002 Dix, T., & Reinhold, D. P. (1991). Chronic and temporary influences on mothers' attributions for children's disobedience. Merrill-Palmer Quar-terly, 37, 251–271. Dixon, W. E., & Smith, P. H. (2003). Who's controlling whom? Infant contributions to maternal play behavior. Infant and Child Development, 12(2), 177–195. doi: 10.1002/icd.283 Dixon, S., Tronick, E., Keefer, C., & Brazelton, T. B. (1981). Motherinfant interaction among the Gusii of Kenya. In T. M. Field, A. M. Sostek, P. Vietze, & P. H. Leiderman (Eds.), Culture and early interactions (pp. 149–170). Hillsdale, NJ: Erlbaum. Doherty, C., Funk, C., Kiley, J., & Weisel, R. (2014). Teaching the children: Sharp ideological differences, some common ground. Retrieved from Pew research center website: http://www.people-press.org/2014/09/18/teaching-the-children-sharp-ideological-differencessome-common-ground/ Domitrovich, C. E., & Bierman, K. L. (2001). Parenting practices and child social adjustment: Multiple pathways of influence. Merrill-Palmer Quarterly, 47(2), 235–263. doi: 10.1353/mpq.2001.0010 Dornbusch, S. M., Ritter, P. L., Leiderman, P. H., Roberts, D. F., & Fraleigh, M. J. (1987). The relation of parenting style to adolescent school performance. Child Development, 58, 1244– 1257. Doty, J. L., Dworkin, J., & Connell, J. H. (2012). Examining digital differences: Parents' online activities. Family Science Review, 17(2), 18–39. Drake, K. L., & Ginsburg, G. S. (2011). Parenting practices of anxious and nonanxious mothers: A multi-method, multi-informant approach. Child & Family Behavior Therapy, 33(4), 299–321. doi: 10.1080/07317107.2011.623101 Draper, P., & Harpending, H. (1988). A sociobiological perspective on the development of human reproductive strategies. In K. B. MacDonald (Ed.), Sociobiological perspectives on human development (pp. 340–372). New York, NY: Springer. Drentea, P., & Moren-Cross, J. L. (2005). Social capital and social support on the web: The

case of an Internet mother site. Sociology of Health & Illness, 27(7), 920–943. doi: 10.1111/j.1467–9566.2005.00464.x Duncan, G. J., & Brooks-Gunn, J. (1997). Consequences of growing up poor. New York, NY: Russell Sage Foundation. Duncan, G. J., Magnuson, K., & Votruba-Drzal, E. (2015). Children and socioeconomic status. In M. H. Bornstein & T. Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R. M. Lerner (Editor-in-Chief), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 534–573). Hoboken, NJ: Wiley. Duncan, L. G., Coatsworth, J. D., & Greenberg, M. T. (2009). A model of mindful parenting: Implications for parent–child relationships and prevention research. Clinical Child and Family Psychology Review, 12(3), 255–270. doi: 10.1007/s10567–009–0046–3 Dunham, P. J., Hurshman, A., Litwin, E., Gusella, J., Ellsworth, C., & Dodd, P. W. (1998). Computer-mediated social support: Single young mothers as a model system. American Journal of Community Psychology, 26(2), 281–306. Dunifon, R. (2012). The influence of grandparents on the lives of children and adolescents. Child Development Perspectives, 7(1), 55–60. doi: 10.1111/cdep.12016 Dunn, J., & Plomin, R. (1990). Separate lives: Why siblings are so different. New York, NY: Basic Books. Dunn, R. R., Davies, J., Harris, N. C., & Gavin, M. C. (2010). Global drivers of human pathogen richness and prevalence. Proceedings of the Royal Society, Biological Sciences, 277, 2587–2595. doi: 10.1098/rspb.2010.0340 Durlak, J. A., & DuPre, E. P. (2008). Implementation matters: A review of research on the influence of implementation on program outcomes and the factors affecting implementation. American Journal of Community Psychology, 41(3–4), 327–350. doi: 10.1007/s10464–008– 9165–0 Dye, J. L. (2010). Fertility of American women: 2008. Current Population Reports, P20–563. Washington, DC: U.S. Census Bureau. Dye, N. S., & Smith, D. B. (1986). Mother love and infant death, 1750–1920. Journal of American History, 73, 329–353. East, P. L., Chien, N. C., & Barber, J. S. (2012). Adolescents' pregnancy intentions, wantedness, and regret: Cross-lagged relations with mental health and harsh parenting. Journal of Marriage and the Family, 74(1), 167–185. doi: 10.1111/j.1741–3737.2011.00885.x Eccles, J. S., & Harold, R. D. (1996). Family involvement in children's and adolescents' schooling. In A. Booth & J. F. Dunn (Eds.), Family-school links: How do they affect educational outcomes? (pp. 3–34). Hillsdale, NJ: Erlbaum.

Edelstein, R. S., Alexander, K., Shaver, P. R., Schaaf, J. M., Quas, J. A., Lovas, G. S., & Goodman, G. S. (2004). Adult attachment style and parental responsiveness during a stressful event. Attachment & Human Development, 6, 31–52. Eibl-Eibesfeldt, I. (1989). Human ethology. New York, NY: Aldine de Gruyter. Eisenberg, N., Vidmar, M. A., Spinrad, T. L., Eggum, N. D., Edwards, A., Gaertner, B., & Kupfer, A. (2010). Mothers' teaching strategies and children's effortful control: A longitudinal study. Developmental Psychology, 46, 1294–1308. Eisengart, S. P., Singer, L. T., Fulton, S., & Baley, J. E. (2003). Coping and psychological Distress in mothers of very low birth weight young children. Parenting: Science and Practice, 3, 49–72. doi: 10.1207/S15327922PAR0301_03 Eisner, M., & Meidert, U. (2011). Stages of parental engagement in a universal parent training program. Journal of Primary Prevention, 32(2), 83–93. doi: 10.1007/s10935–011–0238–8 Ekstrand, G., & Ekstrand, L. H. (1987). Children's perception of norms and sanctions in two cultures. In Ç. Kagitçibasi, Growth and progress in cross-cultural psychology (pp. 171–180). Berwyn, PA: Swets North America. Elder, G. H., Eccles, J. S., Ardelt, M., & Lord, S. (1995). Inner-city parents under economic pressure: Perspective on the strategies of parenting. Journal of Marriage and the Family, 57, 771–784. Elder, G. H., Jr., & Shanahan, M. J. (2006). The life course and human development. In R. E. Lerner (Ed.), Theoretical models of human development (pp. 665–715). In W. Damon (Series Ed.), The handbook of child psychology (6th ed., Vol. 1). Hoboken, NJ: Wiley. doi: 10.1002/9780470147658.chpsy0112 Elek, S. M., Hudson, D. B., & Bouffard, C. (2003). Marital and parenting satisfaction and infant care self-efficacy during the transition to parenthood: The effect of infant sex. Issues in Comprehensive Pediatric Nursing, 26(1), 45–57. doi: 10.1080/01460860390183065 Ellis, B. J., Figueredo, A. J., Brumbach, B. H., & Schlomer, G. L. (2009). The impact of harsh versus unpredictable environments on the evolution and development of life history strategies. Human Nature, 20, 204–268. Ellison, C. G., Bartowski, J. P., & Segal, M. L. (1996). Conservative Protestantism and the parental use of corporal punishment. Social Forces, 74, 1003–1028. Elman, J. L., Bates, E. A., Johnson, M. H., Karmiloff-Smith, A., Parisi, D., & Plunkett, K. (1996). Rethinking innateness: A connectionist perspective on development. Cambridge, MA: MIT Press. Else-Quest, N. M., Hyde, J. S., Goldsmith, H. H., & Van Hulle, C. A. (2006). Gender differences in temperament: A meta-analysis. Psychological Bulletin, 132(1), 33–72. doi:

10.1037/0033–2909.132.1.33 Emmen, R. A. G., Malda, M., Mesman, J., Ekmekci, H., & Van IJzendoorn, M. H. (2012). Sensitive parenting as a cross-cultural ideal: Sensitivity beliefs of Dutch, Moroccan, and Turkish mothers in the Netherlands. Attachment and Human Development, 14, 601–619. doi: 10.1080/14616734.2012.727258 Enkin, M., Keirse, M. J., Chalmers, I., & Enkin, E. (2000). A guide to effective care in pregnancy and childbirth. Oxford, UK: Oxford University Press. Ensor, R., & Hughes, C. (2010). With a little help from my friends: Maternal social support, via parenting, promotes willingness to share in preschoolers born to young mothers. Infant and Child Development, 19(2), 127–141. doi: 10.1002/icd.643 Epstein, J. L., & Sanders, M. G. (2002). Family, school, and community partnerships. In M. H. Bornstein (Ed.), Handbook of parenting: Practical parenting (2nd ed., Vol. 5, pp. 407–437). Mahwah, NJ: Erlbaum. Epstein, R. M., & Hundert, E. M. (2002). Defining and assessing professional competence. Journal of the American Medical Association, 287, 226–235. doi: 10.1001/jama.287.2.226 Erting, C. J., Prezioso, C., & Hynes, M. O. (1994). The interfactional context of deaf mother infant communication. In V. Volterra & C. J. Erting (Eds.), From gesture to language in hearing and deaf children (pp. 97–106). Washington, DC: Gallaudet University Press. Eslinger, P. J. (1996). Conceptualizing, describing, and measuring components of executive function: A summary. In G. R. Lyon & N. A. Krasnegor (Eds.), Attention, memory, and executive function (pp. 367–395). Baltimore, MD: Paul H. Brookes. Eslinger, P. J., Grattan, L. M., Damasio, H., & Damasio, A. R. (1992). Developmental consequences of childhood frontal lobe damage. Archives of Neurology, 49(7), 764–769. Esposito, G., Nakazawa, J., Ogawa, S., Stival, R., Kawashima, A., Putnick, D. L., & Bornstein, M. H. (2014). Baby, you light-up my face: culture-general physiological responses to infants and culture-specific cognitive judgements of adults. PLoS One, 9(10), e106705. doi: 10.1371/journal.pone.0106705 Esposito, G., Valenzi, S., Islam, T., & Bornstein, M. H. (2015). Three physiological responses in fathers and non-fathers' to vocalizations of typically developing infants and infants with Autism Spectrum Disorder. Research in Developmental Disabilities, 43–44, 43–50. Esposito, G., Valenzi, S., Islam, T., Mash, C., & Bornstein, M. H. (2015). Immediate and selective maternal brain responses to own infant faces. Behavioural Brain Research, 278, 40– 43. Evans, G. W. (2004). The environment of childhood poverty. Americal Psychologist, 59, 77. doi: 10.1037/0003–066X.59.2.77

Evans, G. W. (2006). Child development and the physical environment. Annual Reviews of Psychology, 57, 423–451. doi: 10.1146/annurev.psych.57.102904.190057 Evans, G. W., Gonnella, C., Marcynyszyn, L., Gentile, L., & Salpekar, N. (2005). The role of chaos in poverty and children's socioemotional adjustment. Psychological Science, 16, 560– 565. doi: 10.1111/j.0956–7976.2005.01575.x Fabes, R. A., Poulin, R. E., Eisenberg, N., & Madden-Derdich, D. A. (2002). The Coping with Children's Negative Emotions Scale (CCNES): Psychometric properties and relations with children's emotional competence. Marriage & Family Review, 34, 285–310. Fagan, J., & Barnett, M. (2003). The relationship between maternal gatekeeping, paternal competence, mothers' attitudes about the father role, and father involvement. Journal of Family Issues, 24(8), 1020–1043. doi: 10.1177/0192513X03256397 Fang, H., Wang, M., & Xing, X. (2012). Relationship between parenting stress and harsh discipline in preschoolers' parents. Chinese Journal of Clinical Psychology, 20, 835–838. Farnfield, S. (2008). A theoretical model for the comprehensive assessment of parenting. British Journal of Social Work, 38(6), 1076–1099. Farrow, C., & Blissett, J. (2007). The development of maternal self-esteem. Infant Mental Health Journal, 28, 517–535. doi: 10.1037/0012–1649.40.2.234 Fauth, R. C., Leventhal, T., & Brooks-Gunn, J. (2008). Seven years later: Effects of a neighborhood mobility program on poor Black and Latino adults' well-being. Journal of Health and Social Behavior, 49(2), 119–130. doi: 10.1177/002214650804900201 Fearon, R. P., Bakermans-Kranenburg, M. J., Van IJzendoorn, M. H., Lapsley, A. M., & Roisman, G. I. (2010). The significance of insecure attachment and disorganization in the development of children's externalizing behavior: A meta-analytic study. Child Development, 81(2), 435–456. Feinberg, M. E. (2003). The internal structure and ecological context of coparenting: A framework for research and intervention. Parenting: Science and Practice, 3(2), 95–132. doi: 10.1207/S15327922PAR0302_01 Feldman, R. (2007). Parent–infant synchrony and the construction of shared timing; physiological precursors, developmental outcomes, and risk conditions. Journal of Child Psychology and Psychiatry, 48(3–4), 329–354. Feldman, R., Gordon, I., Schneiderman, I., Weisman, O., & Zagoory-Sharon, O. (2010). Natural variations in maternal and paternal care are associated with systematic changes in oxytocin following parent-infant contact. Psychoneuroendocrinology, 35(8), 1133–1141. Feldman, R., Gordon, I., & Zagoory-Sharon, O. (2011). Maternal and paternal plasma, salivary, and urinary oxytocin and parent-infant synchrony: Considering stress and affiliation

components of human bonding. Developmental Science, 14(4), 752–761. doi: 10.1111/j.1467– 7687.2010.01021.x Feldman, R., Granat, A., Pariente, C., Kanety, H., Kuint, J., & Gilboa-Schechtman, E. (2009). Maternal depression and anxiety across the postpartum year and infant social engagement, fear regulation, and stress reactivity. Journal of the American Academy of Child & Adolescent Psychiatry, 48(9), 919–927. doi: 10.1097/CHI.0b013e3181b21651 Feldman, R., Zagoory-Sharon, O., Weisman, O., Schneiderman, I., Gordon, I., Maoz, R.,… Ebstein, R. P. (2012). Sensitive parenting is associated with plasman oxytocin and polymorphisms in the OXTR and CD38 genes. Biological Psychiatry, 72(3), 175–181. Fenning, R. M., Baker, J. K., Baker, B. L., & Crnic, K. A. (2007). Parenting children with borderline intellectual functioning: A unique risk population. Journal Information, 112(2), 107–121. doi: 10.1352/0895–8017(2007)112[107:PCWBIF]2.0.CO;2 Ferguson, C. A. (1977). Baby talk as a simplified register. In C. Snow & C. A. Ferguson (Eds.), Talking to children: Language input and acquisition (pp. 209–235). New York, NY: Cambridge University Press. Fernandez-Duque, E., Valeggia, C. R., & Mendoza, S. P. (2009). The biology of paternal care in human and nonhuman primates. Annual Review of Anthropology, 38, 115–130. doi: 10.1146/annurev-anthro-091908–164334 Festinger, L. (1964). Conflict, decision, and dissonance. Stanford, CA: Stanford University Press. Field, T. M. (2010). Postpartum depression effects on early interactions, parenting, and safety practices: A review. Infant Behavior and Development, 33, 1–6. Fincham, F. D. (1998). Child development and marital relations. Child Development, 69, 543– 574. Finer, L. B., & Henshaw, S. K. (2006). Disparities in rates of unintended pregnancy in the United States, 1994 and 2001. Perspectives on Sexual and Reproductive Health, 38(2), 90– 96. doi: 10.1363/3809006 Finer, L. B., & Zolna, M. R. (2011). Unintended pregnancy in the United States: Incidence and disparities, 2006. Contraception, 84(5), 478–485. Fish, E. W., Shahrokh, D., Bagot, R., Caldji, C., Bredy, T., Szyf, M., & Meaney, M. J. (2004). Epigenetic programming of stress responses through variations in maternal care. Annals of the New York Academy of Sciences, 1036(1), 167–180. doi: 10.1196/annals.1330.011 Fish, M., & Stifter, C. A. (1993). Mother parity as a main and moderating influence on early mother-infant interaction. Journal of Applied Developmental Psychology, 14, 557–572. Fixen, D. L., Naoom, S. D., Blasé, K. A., Friedman, R. M., & Wallace, F. (2005).

Implementation research: A synthesis of the literature. Tampa, FL: Louis d la Parte Florida Mental Health Institute, University of South Florida. Flay, B. R., Biglan, A., Boruch, R. F., Castro, F. G., Gottfredson, D., Kellam, S.,… Ji, P. (2005). Standards of evidence: Criteria for efficacy, effectiveness and dissemination. Prevention Science, 6(3), 151–175. doi: 10.1007/s11121–005–5553-y Fleming, A. S., & Li, M. (2002). Psychobiology of maternal behavior in its early determinants in nonhuman mammals. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 61–97). Mahwah, NJ: Erlbaum. Fletcher, A. C., Darling, N. E., Steinberg, L., & Dornbusch, S. (1995). The company they keep: relation of adolescents' adjustment and behavior to their friends' perceptions of authoritative parenting in the social network. Developmental Psychology, 31(2), 300. Flinn, M. V., Duncan, C. M., Ponzi, D., Quinlan, R. J., Decker, S. A., & Leone, D. V. (2012). Hormones in the wild: Monitoring the endocrinology of family relationships. Parenting: Science and Practice, 12(2–3), 124–133. doi: 10.1080/15295192.2012.683338 Fogel, A., Toda, S., & Kawai, M. (1988). Mother-infant face-to face interaction in Japan and the United States: A laboratory comparison using 3-month-old infants. Developmental Psychology, 24, 398–406. Foley, D. L., Eaves, L. J., Wormly, B., Silberg, J. L., Maes, H. H., Kuhn, J., & Riley, B. (2004). Childhood adversity, monoamine oxidase A genotype, and risk for conduct disorder. Archives of General Psychiatry, 61(7), 738–744. doi: 10.1001/archpsyc.61.7.738 Forgatch, M. S., & DeGarmo, D. S. (1999). Parenting through change: An effective prevention program for single mothers. Journal of Consulting and Clinical Psychology, 67(5), 711–724. Forgatch, M. S., Patterson, G. R., & DeGarmo, D. S. (2005). Evaluating fidelity: Predictive validity for a measure of competent adherence to the Oregon Model of Parent Management Training. Behavior Therapy, 36(1), 3–13. doi: 10.1001/archpsyc.61.7.738 Forman, D. R., O'Hara, M. W., Stuart, S., Gorman, L. L., Larsen, K. E., & Coy, K. C. (2007). Effective treatment for postpartum depression is not sufficient to improve the developing mother-child relationship. Development and Psychopathology, 19(2), 585–602. doi: 10.1017/S0954579407070289 Fraiberg, S., Adelson, E., & Shapiro, V. (1975). Ghosts in the nursery: A psychoanalytic approach to the problems of impaired infant-mother relationships. Journal of the American Academy of Child Psychiatry, 14, 387–421. doi: 10.1016/S0002–7138(09)61442–4 Fraley, R. C. (2002). Attachment stability from infancy to adulthood: Meta-analysis and dynamic modeling of developmental mechanisms. Personality and Social Psychology Review, 6(2), 123–151. doi: 10.1207/S15327957PSPR0602_03

Fraley, R. C., & Shaver, P. R. (2000). Adult romantic attachment: Theoretical developments, emerging controversies, and unanswered questions. Review of General Psychology, 4(2), 132–154. doi: 10.1037/1089–2680.4.2.132 Franco, L. M., Pottick, K. J., & Huang, C. C. (2010). Early parenthood in a community context: Neighborhood conditions, race–ethnicity, and parenting stress. Journal of Community Psychology, 38(5), 574–590. doi: 10.1002/jcop.20382 Frankel, D. G., & Roer-Bornstein, D. (1982). Traditional and modern contributions to changing infant-rearing ideologies of two ethnic communities. Monographs of the Society for Research in Child Development, 4, 1–51. French, V. (2002). History of parenting: The ancient Mediterranean world. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 345– 376). Mahwah, NJ: Erlbaum. Freud, A. (1970). The concept of the rejecting mother. In E. J. Anthony & T. Benedek (Eds.), Parenthood: Its psychology and psychopathology (pp. 376–386). Boston, MA: Little, Brown. (Original work published 1955) Freud, S. (1949). An outline of psycho-analysis. New York, NY: Norton. Friedman, J., & Saunders, N. (2015). Baby care basics. Toronto, Ontario; Robert Rose. Friedman, J., Saunders, N., & Saunders, N. (2013). The A to Z of children's health: A parent's guide from birth to 10 years. Toronto, Ontario: Robert Rose. Fujiwara, T. T., Okuyama, M. M., & Izumi, M. M. (2012). The impact of childhood abuse history, domestic violence and mental health symptoms on parenting behaviour among mothers in Japan. Child: Care, Health and Development, 38(4), 530–537. doi: 10.1111/j.1365– 2214.2011.01272.x Fullgrabe, U. (2002). Psychologie der Eigensicherung: Uberleben ist kein Zufall. Stuttgard, Germany: Boorberg Verlag. Furman, W., & Lanthier, R. (2002). Parenting siblings. In M. H. Bornstein (Ed.), Handbook of parenting (2nd ed., Vol. 1, pp. 165–188). Mahwah, NJ: Erlbaum. Furstenberg, F. F., Jr., Brooks-Gunn, J., & Chase-Lansdale, P. L. (1989). Adolescent fertility and public policy. American Psychologist, 44, 313–320. Gable, S., Crnic, K., & Belsky, J. (1994). Coparenting within the family system: Influences on children's development. Family Relations: Interdisciplinary Journal of Applied Family Studies, 43, 380–386. Galbally, M., Lewis, A. J., van IJzendoorn, M., & Permezel, M. (2011). The role of oxytocin in mother-infant relations: A systematic review of human studies. Harvard Review of Psychiatry, 19(1), 1–14. doi: 10.3109/10673229.2011.549771

Galton, F. (1874). On men of science: Their nature and their nurture. Nature, 9, 344–345. Ganiban, J. M., Saudino, K. J., Ulbricht, J., Neiderhiser, J. M., & Reiss, D. (2008). Stability and change in temperament during adolescence. Journal of Personality and Social Psychology, 95(1), 222. doi: 10.1037/0022–3514.95.1.222 Ganong, L., Coleman, M., & Russell, L. (2015). Children in diverse families. In M. H. Bornstein & T. Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R M. Lerner (Editor-in-Chief), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 133–174). Hoboken, NJ: Wiley. García Coll, C., & Pachter, L. M. (2002). Ethnic and minority parenting. In M. Bornstein (Ed.), Handbook of parenting: Social conditions and applied parenting (2nd ed., Vol. 4). Mahwah, NJ: Erlbaum. Garside, R, & Klimes-Dougan, B. (2002). Socialization of discrete negative emotions: Gender differences and links with psychological distress. Sex Roles, 47(3–4), 115–128. Garrett, P., Ferron, J., Ng'andu, N., Bryant, D., & Harbin, G. (1994). A structural model for the developmental status of young children. Journal of Marriage and the Family, 56(1), 147–163. Gaunt, R. (2008). Maternal gatekeeping: Antecedents and consequences. Journal of Family Issues, 29(3), 373–395. doi: 10.1177/0192513X07307851 Gauthier, A. H., Smeeding, T. M., & Furstenberg, F. F. (2004). Are parents investing less time in children? Trends in selected industrialized countries. Population and Development Review, 30(4), 647–672. doi: 10.1111/j.1728–4457.2004.00036.x Gavin, N. I., Gaynes, B. N., Lohr, K. N., Meltzer-Broder, S., Gartlehner, G., & Swinson, T. (2005). Perinatal depression: A systematic review of prevalence and incidence. Obstetrics & Gynecology, 106(5), 1071–1083. Gaylord, N. K., Kitzmann, K. M., & Coleman, J. K. (2003). Parents' and children's perceptions of parental behavior: Associations with children's psychosocial adjustment in the classroom. Parenting: Science and Practice, 3(1), 23–47. doi: 10.1207/S15327922PAR0301_02 Ge, X., Conger, R. D., Cadoret, R. J., Neiderhiser, J. M., Yates, W., Troughton, E., & Stewart, M. A. (1996). The developmental interface between nature and nurture: A mutual influence model of child antisocial behavior and parent behaviors. Developmental Psychology, 32(4), 574. Geary, D. C. (2000). Evolution and proximate expression of human paternal investment. Psychological Bulletin, 126(1), 55–77. doi: 10.1037/0033–2909.126.1.55 Gecas, V., & Seff, M. A. (1990). Families and adolescents: A review of the 1980s. Journal of Marriage and the Family, 52, 941–958.

George, C., Kaplan, N., & Main, M. (1985). Adult attachment interview. Unpublished manuscript, University of California, Berkeley, CA. Gerdes, A. C., Hoza, B., Arnold, L. E., Pelham, W. E., Swanson, J. M., Wigal, T., & Jensen, P. S. (2007). Maternal depressive symptomatology and parenting behavior: Exploration of possible mediators. Journal of Abnormal Child Psychology, 35(5), 705–714. doi: 10.1007/s10802–007–9134–3 Gershoff, E. T. (2002). Corporal punishment by parents and associated child behaviors and experiences: a meta-analytic and theoretical review. Psychological Bulletin, 128(4), 539. doi: 10.1037/0033–2909.128.4.539 Gervai, J., Nemoda, Z., Lakatos, K., Ronai, Z., Toth, I., Ney, K., & Sasvari-Szekely, M. (2005). Transmission disequilibrium tests confirm the link between DRD4 gene polymorphism and infant attachment. American Journal of Medical Genetics, Part B: Neuropsychiatric Genetics, 132B(1), 126–130. doi: 10.1002/ajmg.b.30102 Gervai, J., Novak, A., Lakatos, K., Toth, I., Danis, I., Ronai, Z.,… Lyons-Ruth, K. (2007). Infant genotype may moderate sensitivity to maternal affective communications: Attachment disorganization, quality of care, and the DRD4 polymorphism. Social Neuroscience, 2(3–4), 307–319. doi: 10.1080/17470910701391893 Gettler, L. T., McDade, T. W., Feranil, A. B., & Kuzawa, C. W. (2011). Longitudinal evidence that fatherhood decreases testosterone in human males. Proceedings of the National Academy of Sciences USA, 108(39), 16194–16199. doi: 10.1073/pnas.1105403108 Gettler, L. T., McDade, T. W., Feranil, A. B., & Kuzawa, C. W. (2012). Prolactin, fatherhood, and reproductive behavior in human males. American Journal of Physical Anthropology, 148(3), 362–370. doi: 10.1002/ajpa.22058 Gettler, L. T., McDade, T. W., & Kuzawa, C. W. (2011). Cortisol and testosterone in Filipino young adult men: Evidence for co-regulation of both hormones by fatherhood and relationship status. American Journal of Human Biology, 23(5), 609–620. doi: 10.1002/ajhb.21187 Gibson, E., Dembofsky, C. A., Rubin, S., & Greenspan, J. S. (2000). Infant sleep position practices 2 years into the “back to sleep” campaign. Clinical Pediatrics, 39(5), 285–289. doi: 10.1177/000992280003900505 Global Initiative to End All Corporal Punishment of Children. (2013). Ending legalised violence against children. Retrieved from http://www.endcorporalpunishment.org/pages/pdfs/reports/GlobalReport2013.pdf Glocker, M. L., Langleben, D. D., Ruparel, K., Loughead, J. W., Gur, R. C., & Sachser, N. (2009a). Baby schema in infant faces induces cuteness perception and motivation for caretaking in adults. Ethology, 115(3), 257–263. doi: 10.1111/j.1439–0310.2008.01603.x Glocker, M. L., Langleben, D. D., Ruparel, K., Loughead, J. W., Valdez, J. N., Griffin, M. D.,

… Gur, R. C. (2009b). Baby schema modulates the brain reward system in nulliparous women. Proceedings of the National Academy of Sciences USA, 106(22), 9115–9119. doi: 10.1073/pnas.0811620106 Golan, S., Spiker, D., Peterson, D., Mercier, B., Snow, M., & Williamson, C. (2008). Parent voices: A statewide look. Washington State Department of Early Learning Parent Needs Assessment: Phone survey (SRI Project 18252). Menlo Park, CA: SRI International. Goldberg, S. (1977). Infant development and mother-infant interaction in urban Zambia. In P. H. Leiderman, S. R. Tulkin, & A. Rosenfeld (Eds.), Culture and infancy: Variations in the human experience (pp. 211–243). New York, NY: Academic Press. Goldberg, S., & DiVitto, B. (2002). Parenting children born preterm. In M. H. Bornstein (Ed.), Handbook of parenting: Children and parenting (2nd ed., Vol. 1, pp. 329–354). Mahwah, NJ: Erlbaum. Golden, T. (2012). Altering the effects of work and family conflict on exhaustion: Telework during traditional and nontraditional work hours. Journal of Business Psychology, 27, 255– 269. Golombok, S. (2002). Parenting and contemporary reproductive technologies. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 339–360). Mahwah, NJ: Erlbaum. Golombok, S. (2013). Families created by reproductive donation: Issues and research. Child Development Perspectives, 7(1), 61–65. doi: 10.1111/cdep.12015 Gonzalez, A., Lovic, V., Ward, G. R., Wainwright, P. E., & Fleming, A. S. (2000). Intergenerational effects of complete maternal deprivation and replacement stimulation on maternal behavior and emotionality in female rats. Developmental Psychobiology, 38(1), 11– 32. doi: 10.1002/1098–2302(2001)38:13.0.CO;2-B Goodman, J. H. (2004). Paternal postpartum depression, its relationship to maternal postpartum depression, and implications for family health. Journal of Advanced Nursing, 45(1), 26–35. doi: 10.1046/j.1365–2648.2003.02857.x Goodnow, J. J. (2002). Parents' knowledge and expectations: Using what we know. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 439–460). Mahwah, NJ: Erlbaum. Goodnow, J. J., Cashmore, R., Cotton, S., & Knight, R. (1984). Mothers' developmental timetables in two cultural groups. International Journal of Psychology, 19, 193–205. Goodnow, J. J., & Collins, W. A. (1990). Development according to parents: The nature, sources, and consequences of parents' ideas. Hillsdale, NJ: Erlbaum. Goodnow, J. J., & Lawrence, J. A. (2015). Children and cultural context. In M. H. Bornstein &

T. Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R. M. Lerner (Editor-in-Chief), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 746–786). Hoboken, NJ: Wiley. Goodson, B. D., Layzer, J. I., St. Pierre, R. G., Bernstein, L. S., & Lopez, M. (2000). Effectiveness of a comprehensive, five-year family support program for low-income children and their families: Findings from the Comprehensive Child Development Program. Early Childhood Research Quarterly, 15, 5–39. doi: 10.1016/S0885–2006(99)00040-X Gottfredson, D., Kumpfer, K., Polizzi-Fox, D., Wilson, D., Puryear, V., Beatty, P., & Vilmenay, M. (2006). The Strengthening Washington DC Families Project: A randomized effectiveness trial of family-based prevention. Prevention Science, 7(1), 57–74. doi: 10.1007/s11121–005– 0017-y Gottfried, A. E., Gottfried, A. W., & Bathurst, K. (2002). Maternal and dual-earner employment status and parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 207–229). Mahwah, NJ: Erlbaum. Gottlieb, G. (1995). Some conceptual deficiencies in “developmental” behavior genetics. Human Development, 38(3), 131–141. Gottman, J. M., & Levenson, R. W. (2000). The timing of divorce: Predicting when a couple will divorce over a 14-year period. Journal of Marriage and Family, 62(3), 737–745. doi: 10.1111/j.1741–3737.2000.00737.x Grattan, L. M., & Eslinger, P. J. (1992). Long-term psychological consequences of childhood frontal lobe lesion in patient DT. Brain and Cognition, 20(1), 185–195. Green, R. (1954). Employment counseling for the hard of hearing. Volta Review, 56, 209–212. Greenberger, E., & Goldberg, W. A. (1989). Work, parenting, and the socialization of children. Developmental Psychology, 25(1), 22–35. Greene, M. L., & Way, N. (2005). Self-esteem trajectories among ethnic minority adolescents: A growth curve analysis of the patterns and predictors of change. Journal of Research on Adolescence, 15(2), 151–178. Greenfield, P. M. (2009). Linking social change and developmental change: shifting pathways of human development. Developmental Psychology, 45(2), 401. Greenfield, P. M., Keller, H., Fuligni, A., & Maynard, A. (2003). Cultural pathways through universal development. Annual Review of Psychology, 54(1), 461–490. doi: 10.1146/annurev.psych.54.101601.145221 Groh, A. M., Roisman, G. I., van IJzendoorn, M. H., Bakermans-Kranenburg, M. J., & Fearnon, R. P. (2012). The significance of insecure and disorganized attachment for children's internalizing symptoms: A meta-analytic study. Child Development, 83(2), 591–610. doi:

10.1111/j.1467–8624.2011.01711.x Grossman, F. K., Eichler, L. W., & Winikoff, S. A. (1980). Pregnancy, birth and parenthood. San Francisco, CA: Jossey-Bass. Grusec, J. E., & Goodnow, J. J. (1994). Impact of parental discipline methods on the child's internalization of values: A reconceptualization of current points of view. Developmental Psychology, 30(1), 4–19. Grusec, J. E., & Kuczynski, L. (1980). Direction of effect in socialization: A comparison of the parent vs. the child's behavior as determinants of disciplinary techniques. Developmental Psychology, 16, 1–9. Grussu, P., Quatraro, R. M., & Nasta, M. T. (2005). Profile of mood states and parental attitudes in motherhood: Comparing women with planned and unplanned pregnancies. Birth, 32(2), 107–14. doi: 10.1111/j.0730–7659.2005.00353.x Grych, J. H. (2002). Marital relationships and parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Applied parenting (2nd ed., Vol. 4, pp. 203–225). Mahwah, NJ: Erlbaum. Guryan, J., Hurst, E., & Kearney, M. S. (2008). Parental education and parental time with children. Journal of Economic Perspectives, 22(3), 23–46. doi: 10.3386/w13993 Guterman, N. B., Lee, S. J., Taylor, C. A., & Rathouz, P. J. (2009). Parental perceptions of neighborhood processes, stress, personal control, and risk for physical child abuse and neglect. Child Abuse & Neglect, 33(12), 897–906. doi: 10.1016/j.chiabu.2009.09.008 Haltigan, J. D., Leerkes, E. M., Wong, M. S., Fortuna, K., Roisman, G. I., Supple, A.,… Plamondon, A. (2014). Adult attachment states of mind: Measurement invariance across ethnicity and associations with maternal sensitivity. Child Development, 85(3), 1019–1035. doi: 10.1111/cdev.12180 Haltigan, J. D., Roisman, G. I., & Fraley, R. C. (2013). The predictive significance of early caregiving experiences for symptoms of psychopathology through midadolescence: Enduring or transient effects? Development and Psychopathology, 25(1), 209–221. doi: 10.1017/S0954579412000260 Hamm, J. V. (2001). Barriers and bridges to positive cross-ethnic relations African American and White parent socialization beliefs and practices. Youth & Society, 33(1), 62–98. doi: 10.1177/0044118X01033001003 Harlow, H. F. (1958). The nature of love. American Psychologist, 13(12), 673–685. Harris, C. C. (1983). The family and industrial society. London, UK: Allen & Unwin. Harris, J. R. (1998). The nurture assumption. New York, NY: Free Press. Hart, B., & Risley, T. R. (1995). Meaningful differences in the everyday experience of young

American children. Baltimore, MD: Paul H. Brookes. Harwood, R. L. (1992). The influence of culturally derived values on Anglo and Puerto Rican mothers' perceptions of attachment behavior. Child Development, 63, 822–839. Harwood, R., Leyendecker, B., Carlson, V., Asencio, M., & Miller, A. (2002). Parenting among Latino families in the U.S. In M. H. Bornstein (Ed.), Handbook of parenting: Applied parenting (2nd ed., Vol. 4, pp. 21–46). Mahwah, NJ: Erlbaum. Harwood, R. L., Miller, J. G., & Irizarry, N. L. (1995). Culture and attachment: Perceptions of the child in context. New York, NY: Guilford Press. Hashima, P. Y., & Amato, P. R. (1994). Poverty, social support, and parental behavior. Child Development, 65, 394–403. Haugaard, J., & Hazan, C. (2002). Foster parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Children and parenting (2nd ed., Vol. 1, pp. 313–327). Mahwah, NJ: Erlbaum. Hebbeler, K. M., & Gerlach-Downie, S. G. (2002). Inside the black box of home visiting: A qualitative analysis of why intended outcomes were not achieved. Early Childhood Research Quarterly, 17(1), 28–51. doi: 10.1016/S0885–2006(02)00128-X Henrich, J., Heine, S. J., & Norenzayan, A. (2010). The weirdest people in the world? Behavioral and Brain Sciences, 33(2–3), 61–135. doi: 10.1017/S0140525X0999152X Heron, J., O'Connor, T. G., Evans, J., Golding, J., Glover, V., & Team, A. S. (2004). The course of anxiety and depression through pregnancy and the postpartum in a community sample. Journal of Affective Disorders, 80(1), 65–73. doi: 10.1016/j.jad.2003.08.004 Hess, C. R., Teti, D. M., & Hussey-Gardner, B. (2004). Self-efficacy and parenting of highrisk infants: The moderating role of parent knowledge of infant development. Journal of Applied Developmental Psychology, 25(4), 423–437. doi: 10.1016/j.appdev.2004.06.002 Hetherington, E. M., Henderson, S. H., Reiss, D., Anderson, E. R., Bridges, M., Chan, R. W., … & Taylor, L. C. (1999). Adolescent siblings in stepfamilies: Family functioning and adolescent adjustment. Monographs of the Society for Research in Child Development, 64(4), 222. Hetherington, E. M., & Stanley-Hagan, M. M. (2002). Parenting in divorced and remarried families. In M. H. Bornstein (Ed.), Handbook of parenting. Status and social conditions of parenting (2nd ed., Vol. 3, pp. 287–315). Mahwah, NJ: Erlbaum. Hewlett, B. S., Lamb, M. E., Shannon, D., Leyendecker, B., & Schölmerich, A. (1998). Culture and early infancy among central African foragers and farmers. Developmental Psychology, 34(4), 653. Hickson, G. B., & Clayton, E. W. (2002). Parents and their children's doctors. In M. H. Bornstein (Ed.), Handbook of parenting: Practical parenting (Vol. 5, pp. 439–462). Mahwah,

NJ: Erlbaum. Ho, D. Y. F. (1994). Cognitive socialization in Confucian heritage cultures. In P. M. Greenfield & R. R. Cocking (Eds.), Cross-cultural roots of minority child development (pp. 285–313). Hillsdale, NJ: Erlbaum. Hodapp, R. M. (2002). Parenting children with mental retardation. In M. H. Bornstein (Ed.), Handbook of parenting: Children and parenting (2nd ed., Vol. 1, pp. 355–381). Mahwah, NJ: Erlbaum. Hodapp, R. M., & Ly, T. M. (2005). Parenting children with developmental disabilities. In T. Luster & L. Okagaki (Eds.), Parenting: An ecological perspective (2nd ed., pp. 177–201). Mahwah, NJ: Erlbaum. Hoek, H. W., Brown, A. S., & Susser, E. S. (1999). The Dutch Famine Studies: Prenatal nutritional deficiency and schizophrenia. In E. Susser, A. Brown, & J. Gorman (Eds.), Prenatal exposures in schizophrenia (pp. 135–161). Washington, DC: American Psychiatric Press, Inc. Hoff, E. (2003). Causes and consequences of SES-related differences in parent-to-child speech. In M. H. Bornstein & R. H. Bradley (Eds.), Socioeconomic status, parenting, and child development (pp. 147–160). Mahwah, NJ: Erlbaum. Hoff, E., Laursen, B., & Tardif, T. (2002). Socioeconomic status and parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 231–252). Mahwah, NJ: Erlbaum. Hofferth, S. L. (1987). The children of teen childbearers. In S. L. Hofferth & C. D. Hayes (Eds.), Risking the future: Adolescent sexuality, pregnancy and childbearing (pp. 174–206). Washington, DC: National Academies Press. Hofferth, S. L., & Anderson, K. G. (2003). Are all dads equal? Biology versus marriage as a basis for paternal investment. Journal of Marriage and the Family, 65(1), 213–232. doi: 10.1111/j.1741–3737.2003.00213.x Hofferth, S. L., Pleck, J. H., & Vesely, C. K. (2012). The transmission of parenting from fathers to sons. Parenting: Science and Practice, 12(4), 282–305. doi: 10.1080/15295192.2012.709153 Hoffman, L. W. (1991). The influence of the family environment on personality: Accounting for sibling differences. Psychological Bulletin, 110, 187–203. Holden, G. W. (1997). Parents and the dynamics of child rearing. Boulder, CO: Westview Press. Holden, G. W. (2002). Perspectives on the effects of corporal punishment: Comment on Gershoff. Psychological Bulletin, 128(4), 590–595. doi: 10.1037/0033–2909.128.4.590

Holden, G. W. (2009). Parenting: A dynamic perspective. Thousand Oaks, CA: Sage. Holden, G. W., & Buck, M. J. (2002). Parental attitudes toward childrearing. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 537–562). Mahwah, NJ: Erlbaum. Holden, G. W., Coleman, S., & Schmidt, K. L. (1995). Why 3-year-old children get spanked: Parent and child determinants in a sample of college-educated mothers. Merrill-Palmer Quarterly, 4I, 431–452. Holden, G. W., & Miller, P. C. (1999). Enduring and different: A meta-analysis of the similarity in parents' child rearing. Psychological Bulletin, 125(2), 223–254. Holden, G. W., Thompson, L., Zambarano, R., & Marshall, L. (1997). Child effects as a source of change in maternal attitudes toward corporal punishment. Journal of Social and Personal Relationships, I4, 481–490. Hoover-Dempsey, K. V., & Sandler, H. M. (1997). Why do parents become involved in their children's education? Review of Educational Research, 67, 3–42. Hopkins, B., & Westra, T. (1989). Maternal expectations of their infants' development: Some cultural differences. Developmental Medicine and Child Neurology, 31(3), 384–390. Hopkins, B., & Westra, T. (1990). Motor development, maternal expectations, and the role of handling. Infant Behavior and Development, 13(1), 117–122. Houck, C. D., Rodrigue, J. R., & Lobato, D. (2007). Parent–adolescent communication and psychological symptoms among adolescents with chronically ill parents. Journal of Pediatric Psychology, 32(5), 596–604. doi: 10.1093/jpepsy/jsl048 Howard, K. S., & Brooks-Gunn, J. (2009). Relationship supportiveness during the transition to parenting among married and unmarried parents. Parenting: Science and Practice, 9, 123– 142. doi: 10.1080/15295190802656828 Howes, C., & Oldman, E. (2001). Processes in the formation of attachment relationships with alternative caregivers. In A. Göncü & E. L. Klein (Eds.), Children in play, story, and school (pp. 267–287). New York, NY: Guilford Press. Howes, C., & Spieker, S. (2008). Attachment relationships in the context of multiple caregivers. In J. Cassidy & P. R. Shaver (Eds.), Handbook of attachment: Theory, research, and clinical applications (2nd ed., pp. 317–332). New York, NY: Guilford Press. Hrdy, S. B. (1999). Mother nature: A history of mothers, infants, and natural selection. New York, NY: Pantheon. Hrdy, S. B. (2009). Mothers and others: The evolutionary origins of mutual understanding. Cambridge, MA: Harvard University Press.

Hsueh, J., & Yoshikawa, H. (2007). Working nonstandard schedules and variable shifts in low income families: Associations with parental psychological well-being, family functioning, and child well-being. Developmental Psychology, 43(3), 620. doi: 10.1037/0012–1649.43.3.620 Huang, K. Y., O'Brien Caughy, M., Genevro, J. L., & Miller, T. L. (2005). Maternal knowledge of child development and quality of parenting among White, African-American, and Hispanic mothers. Applied Developmental Psychology, 26(2), 149–170. doi: 10.1016/j.appdev.2004.12.001 Huang, L, Malone, P. S., Lansford, J. E., Deater-Deckard, K., Di Giunta, L., Bombi, A. S.,… Bacchini, D. (2011). Measurement invariance of mother reports of discipline in different cultural contexts. Family Science, 2(3), 212–219. doi: 10.1080/19424620.2011.655997 Hudson, D. B., Campbell-Grossman, C., Elek, S., Fleck, M., & Shipman, A. (2003). Effects of the new parents network on first-time fathers' parenting satisfaction during the transition to parenthood. Comprehensive Pediatric Nursing, 26, 217–229. Huh, D., Tristan, J., Wade, E., & Stice, E. (2006). Does problem behavior elicit poor parenting? A prospective study of adolescent girls. Journal of Adolescent Research, 21(2), 185–204. doi: 10.1177/0743558405285462 Hulbert, A. (2003). Raising America: Experts, parents, and a century of advice about children. New York, NY: Knopf. Hurley, K. D., Huscroft-D'Angelo, J., Trout, A., Griffith, A., & Epstein, M. (2014). Assessing parenting skills and attitudes: A review of the psychometrics of parenting measures. Journal of Child and Family Studies, 23(5), 812–823. doi: 10.1007/s10826–013–9733–2 Huws, J. C., Jones, R. S., & Ingledew, D. K. (2001). Parents of children with autism using an email group: A grounded theory study. Journal of Health Psychology, 6(5), 569–584. doi: 10.1177/135910530100600509 Hyde, J. S. (2005). The gender similarities hypothesis. American Psychologist, 60, 581–592. Hyde, J. S. (2014). Gender similarities and differences. Annual Review of Psychology, 65, 373–398. Inglehart, R. (1997). Modernization and postmodernization: Cultural, economic, and political change in 43 societies (Vol. 19). Princeton, NJ: Princeton University Press. Inglehart, R., & Baker, W. E. (2000). Modernization, cultural change, and the persistence of traditional values. American Sociological Review, 65, 19–51. Insel, T. R. (2003). Is social attachment an addictive disorder? Physiology and Behavior, 79(3), 351–357. doi: 10.1016/S0031–9384(03)00148–3 Ispa, J., Fine, M., Halgunseth, L., Harper, S., Robinson, J., & Boyce, L. (2004). Maternal intrusiveness, maternal warmth, and mother-toddler relationship outcomes: Variations across

low-income ethnic and acculturation groups. Child Development, 75(6), 1613–1631. doi: 10.1111/j.1467–8624.2004.00806.x Izzo, C., Weiss, L., Shanahan, T., & Rodriguez-Brown, F. (2000). Parental self-efficacy and social support as predictors of parenting practices and children's socioemotional adjustment in Mexican immigrant families. Journal of Prevention & Intervention in the Community, 20(1– 2), 197–213. doi: 10.1300/J005v20n01_13 Jacobson, J. L., Boersma, D. C., Fields, R. B., & Olson, K. L. (1983). Paralinguistic features of adult speech to infants and small children. Child Development, 54(2), 436–442. Jaffee, S. R., Belsky, J., Harrington, H., Caspi, A., & Moffitt, T. E. (2006). When parents have a history of conduct disorder: How is the caregiving environment affected? Journal of Abnormal Psychology, 115(2), 309–319. doi: 10.1037/0021–843X.115.2.309 Jaffee, S. R., Caspi, A., Moffitt, T. E., & Taylor, A. (2004). Physical maltreatment victim to antisocial child: Evidence of an environmentally mediated process. Journal of Abnormal Psychology, 113(1), 44. doi: 10.1037/0021–843X.113.1.44 Jakobson, R. (1969). Kindersprache, aphasie und allgemeine lautgesetze (Vol. 330). Frankfurt, Germany: Suhrkamp. (Original work published in 1941.) Jambunathan, S., & Counselman, K. (2002). Parenting attitudes of Asian Indian mothers living in the United States and in India. Early Child Development and Care, 172(6), 657–662. doi: 10.1080/03004430215102 Jang, S. J., & Smith, C. A. (1997). A test of reciprocal causal relationships among parental supervision, affective ties, and delinquency. Journal of Research in Crime and Delinquency, 34(3), 307–336. Jarrett, P. G. (1998). Logistics in the health care industry. International Journal of Physical Distribution & Logistics Management, 28(9/10), 741–772. Jenkins, J. M., Rasbash, J., & O'Connor, T. G. (2003). The role of the shared family context in differential parenting. Developmental Psychology, 39, 99–113. Jensen, C. J., & Champagne, F. A. (2012). Epigenetic and neurodevelopmental perspectives on variation in parenting behavior. Parenting: Science and Practice, 12(2–3), 202–211. doi: 10.1080/15295192.2012.683358 Johnston, C., & Mash, E. (1989). A measure of parenting satisfaction and efficacy. Journal of Clinical Child Psychology, 18, 167–175. Joosen, K. J., Mesman, J., Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2012). Maternal sensitivity to infants in various settings predicts harsh discipline in toddlerhood. Attachment & Human Development, 14, 101–117. Joshi, A., & Gutierrez, B. J. (2006). Parenting stress in parents of Hispanic adolescents. North

American Journal of Psychology, 8(2), 209–216. Juffer, F., Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2007). Promoting positive parenting. Mahwah, NJ: Erlbaum. Kaffman, A., & Meaney, M. J. (2007). Neurodevelopmental sequelae of postnatal maternal care in rodents: Clinical and research implications of molecular insights. Journal of Child Psychology and Psychiatry, 48(3–4), 224–244. doi: 10.1111/j.1469–7610.2007.01730.x Kagan, J., & Moss, H. A. (1962). Birth to maturity: A study in psychological development. New York: Wiley. Kaminski, J. W., Valle, L. A., Filene, J. H., & Boyle, C. L. (2008). A meta-analytic review of components associated with parent training program effectiveness. Journal of Abnormal Child Psychology, 36(4), 567–589. doi: 10.1007/s10802–007–9201–9 Karraker, K. H., & Coleman, P. K. (2005). The effects of child characteristics on parenting. In T. Luster & L. Okagaki (Eds.), Parenting: An ecological perspective (2nd ed., pp. 147–176). Mahwah, NJ: Erlbaum. Karraker, K. H., & Stern, M. (1990). Infant physical attractiveness and facial expression: Effects on adult perceptions. Basic and Applied Social Psychology, 11(4), 371–385. Kashima, Y., Bain, P., Haslam, N., Peters, K., Laham, S., Whelan, J.,… Fernando, J. (2009). Folk theory of social change. Asian Journal of Social Psychology, 12, 227–246. doi: 10.1111/j.1467–839X.2009.01288.x Kaufman, J., Yang, B. Z., Douglas-Palumberi, H., Houshyar, S., Lipschitz, D., Krystal, J., & Gelernter, J. (2004). Social supports and serotonin transporter gene moderate depression in maltreated children. Proceedings of the National Academy of Sciences USA, 101(49), 17316– 17321. doi: 10.1073/pnas.0404376101 Keen, C. L., Bendich, A., & Willhite, C. C. (1993). Maternal nutrition and pregnancy outcome. Annals of the New York Academy of Sciences (USA), 678, 1–372. Keller, H., Kärtner, J., Borke, J., Yovski, R., & Kleis, A. (2005). Parenting styles and the development of the categorical self: A longitudinal study on mirror self-recognition in Cameroonian Nso and German families. International Journal of Behavioral Development, 29, 496–504. Keller, H., Lamm, B., Abels, M., Yovsi, R., Borke, J., Jensen, H.,… Chaudhary, N. (2006). Cultural models, socialization goals, and parenting ethnotheories: A multicultural analysis. Journal of Cross-Cultural Psychology, 37(2), 155–172. doi: 10.1177/0022022105284494 Keller, H., Lohaus, A., Kuensemueller, P., Abels, M., Yovsi, R., Voelker, S., Jensen, H.,… Mohite, P. (2004). The bio-culture of parenting: Evidence from five cultural communities. Parenting: Science and Practice, 4, 25–50. doi: 10.1207/s15327922par0401_2

Kendler, K. S., & Baker, J. H. (2007). Genetic influences on measures of the environment: A systematic review. Psychological Medicine, 37(5), 615–626. doi: 10.1017/S0033291706009524 Kenkel, W. M., Paredes, J., Yee, J. R., Pournajafi-Nazarloo, H., Bales, K. L., & Carter, C. S. (2012). Neuroendocrine and behavioural responses to exposure to an infant in male prairie voles. Journal of Neuroendocrinology, 24(6), 874–886. doi: 10.1111/j.1365– 2826.2012.02301.x Kennedy, A. E., Rubin, K. H., Hastings, P. D., & Maisel, B. (2004). Longitudinal relations between child vagal tone and parenting behavior: 2 to 4 years. Developmental Psychobiology, 45, 10–21. Kerr, D., Capaldi, D., Pears, K., & Owen, L. (2009). A prospective three-generational study of fathers' constructive parenting: Influences from family of origin, adolescent adjustment, and offspring temperament. Developmental Psychology, 45(5), 1257–1275. doi: 10.1037/a0015863 Kerr, M., & Stattin, H. (2000). What parents know, how they know it, and several forms of adolescent adjustment: further support for a reinterpretation of monitoring. Developmental Psychology, 36(3), 366. doi: 10.1037/0012–1649.36.3.366 Kerr, M., & Stattin, H. (2003). Parenting of adolescents: Action or reaction? In A. Booth & A. Crouter (Eds.), Children's influence on family dynamics: The neglected side of family relationships (pp. 121–151). Mahwah, NJ: Erlbaum. Kerr, M., Stattin, H., Biesecker, G., & Ferrer-Wreder, L. (2003). Relationships with parents and peers in adolescence. In I. B. Weiner (Editor-in-Chief), Handbook of psychology (pp. 395–419). Hoboken, NJ: Wiley. doi: 10.1002/0471264385.wei0616 Kerr, M., Stattin, H., & Trost, K. (1999). To know you is to trust you: Parents' trust is rooted in child disclosure of information. Journal of Adolescence, 22(6), 737–752. Kertz, S. J., Smith, C. L., Chapman, L. K., & Woodruff-Borden, J. (2008). Maternal sensitivity and anxiety: Impacts on child outcome. Child & Family Behavior Therapy, 30(2), 153–171. doi: 10.1080/07317100802060336 Kim, J., & Wickrama, K. A. S. (2013). Mothers' working status and infant development: Mediational processes. Journal of Family Issues, 35(11), 1473–1496. Advance online publication. doi: 10.1177/0192513X13496414 Kim, P., Leckman, J. F., Mayes, L. C., Feldman, R., Wang, X., & Swain, J. E. (2010a). The plasticity of human maternal brain: Longitudinal changes in brain anatomy during the early postpartum period. Behavioral Neuroscience, 124(5), 695–700. Kim, P., Leckman, J. F., Mayes, L. C., Newman, M. A., Feldman, R., & Swain, J. E. (2010b). Perceived quality of maternal care in childhood and structure and function of mothers' brain.

Developmental Science, 13(4), 662–673. doi: 10.1111/j.1467–7687.2009.00923.x Kim-Cohen, J., Caspi, A., Rutter, M., Tomás, M., & Moffitt, T. (2006a). The caregiving environments provided to children by depressed mothers with or without an antisocial history. American Journal of Psychiatry, 163(6), 1009 –1018. Kim-Cohen, J., Caspi, A., Taylor, A., Williams, B., Newcombe, R., Craig, I. W., & Moffitt, T. E. (2006b). MAOA, maltreatment, and gene-environment interaction predicting children's mental health: New evidence and a meta-analysis. Molecular Psychiatry, 11(10), 903–913. King, P. E., & Boyatzis, C. J. (2015). Religious and spiritual development. In M. E. Lamb (Ed.), Socioemotional processes: Volume 3 of the Handbook of child psychology and developmental science (7th ed., pp. 975–1021). Editor-in-chief: R. M. Lerner. Hoboken, NJ: Wiley. King, V., & Elder, G. H., Jr. (1998). Perceived self-efficacy and grandparenting. Journals of Gerontology Series B: Psychological Sciences and Social Sciences, 53B, S249–S257. Kinlaw, R., Kurtz-Costes, B., & Goldman-Fraser, J. (2001). Mothers' achievement beliefs and behaviors and their children's school readiness: A cultural comparison. Journal of Applied Developmental Psychology, 22(5), 493–506. doi: 10.1016/S0193–3973(01)00090–9 Kirby, A., Edwards, L., & Hughes, A. (2008). Parents' concerns about children with specific learning difficulties: Insights gained from an online message centre. Support for Learning, 23(4), 193–200. Klahr, A. M., & Burt, S. A. (2014). Elucidating the etiology of individual differences in parenting: A meta-analysis of behavioral genetic research. Psychology Bulletin, 140(2), 544– 586. doi: 10.1037/a0034205 Klein, P. (1988). Stability and change in interaction of Israeli mothers and infants. Infant Behavior and Development, 11, 55–70. Knafo, A., & Jaffee, S. R. (2013). Gene–environment correlation in developmental psychopathology. Development and Psychopathology, 25(01), 1–6. doi: 10.1017/S0954579412000855 Kochanska, G. (1995). Children's temperament, mothers' discipline, and security of attachment: Multiple pathways to emerging internalization. Child Development, 66, 597–615. Kochanska, G. (1997). Multiple pathways to conscience for children with different temperaments: From toddlerhood to age five. Developmental Psychology, 33, 228–240. Kochanska, G., & Aksan, N. (2004). Development of mutual responsiveness between parents and their young children. Child Development, 75(6), 1657–1676. doi: 10.1111/j.14678624.2004.00808.x Kochanska, G., Clark, L., & Goldman, M. (1997). Implications of mothers' personality for

parenting and their young children's developmental outcomes. Journal of Personality, 65, 389–420. Koenig, J. L., Barry, R. A., & Kochanska, G. (2010). Rearing difficult children: Parents' personality and children's proneness to anger as predictors of future parenting. Parenting: Science and Practice, 10, 258–273. doi: 10.1080/15295192.2010.492038 Kohen, D. E., Leventhal, T., Dahinten, V. S., & McIntosh, C. N. (2008). Neighborhood disadvantage: Pathways of effects for young children. Child Development, 79(1), 156–169. doi: 10.1111/j.1467–8624.2007.01117.x Kohn, M. L. (1963). Social class and parent-child relationships: An interpretation. American Journal of Sociology, 68, 471–480. Kohn, M. L. (1969). Class and conformity: A study in values. Oxford, UK: Dorsey. Kohn, M. L. (1979). The effects of social class on parental values and practices. In D. Reiss & H. A. Hoffman (Eds.), The American family: Dying or developing (pp. 45–68). New York, NY: Plenum Press. Kong, A., Frigge, M. L., Masson, G., Besenbacher, S., Sulem, P., Magnusson, G.,… Stefansson, K. (2012). Rate of de novo mutations and the importance of father's age to disease risk. Nature, 488(7412), 471–475. doi: 10.1038/nature11396 Konner, M. (1977). Infancy among the Kalahari Desert San. In P. H. Leiderman, S. R. Tulkin, & A. Rosenfeld (Eds.), Culture and infancy: Variations in the human experience (pp. 287– 328). San Diego, CA: Academic Press. Konner, M. J. (1991). Universals of behavioral development in relation to brain myelination. In A. C. Petersen (Ed.), Brain maturation and cognitive development. New York, NY: Aldine de Gruyter. Konner, M. J. (2010). The evolution of childhood: Relationships, emotion, mind. Cambridge, MA: Harvard University Press. Korfmacher, J., Green, B., Staerkel, F., Peterson, C., Cook, G., Roggman, L.,… Schiffman, R. (2008). Parent involvement in early childhood home visiting. Child & Youth Care Forum, 37(4), 171–196. doi: 10.1007/s10566–008–9057–3 Korfmacher, J., O'Brien, R., Hiatt, S., & Olds, D. (1999). Differences in program implementation between nurses and paraprofessionals providing home visits during pregnancy and infancy: A randomized trial. American Journal of Public Health, 89, 1847–1851. Kornhaber, A. (2002). The grandparent's guide: The definitive guide to coping with the challenges of modern grandparenting. New York, NY: McGraw Hill Professional. Kotov, R., Gamez, W., Schmidt, F., & Watson, D. (2010). Linking “big” personality traits to anxiety, depressive, and substance use disorders: A meta-analysis. Psychological Bulletin,

136(5), 768–821. doi: 10.1037/a0020327 Kovan, N. M., Chung, A. L., & Sroufe, L. A. (2009). The intergenerational continuity of observed early parenting: A prospective, longitudinal study. Developmental Psychology, 45(5), 1205–1213. doi: 10.1037/a0016542 Kozorovitskiy, Y., Hughes, M., Lee, K., & Gould, E. (2006). Fatherhood affects dendritic spines and vasopressin V1a receptors in the primate prefrontal cortex. Nature Neuroscience, 9(9), 1094–1095. doi: 10.1038/nn1753 Krapf, M., Ursprung, H. W., & Zimmermann, C. (2014). Parenthood and productivity of highly skilled labor: Evidence from the groves of Academe Research Division. Working Paper No. 2014–001A. Retrieved from Federal Reserve Bank of St. Louis Working Paper Series website: http://research.stlouisfed.org/wp/2014/2014–001.pdf Kreider, R. M., & Ellis, R. (2011). Living arrangements of children: 2009. Current Population Reports, P70–126. Washington, DC: U.S. Census Bureau. Krueger, A., Kahneman, D., Schkade, D., Schwarz, N., & Stone, A. (2009). National time accounting: The currency of life. In A. Krueger (Ed.), National time accounting and subjective well-being (pp. 9–84). Chicago, IL: University of Chicago Press. Kuhn, J. C., & Carter, A. S. (2006). Maternal self-efficacy and associated parenting cognitions among mothers of children with autism. American Journal of Orthopsychiatry, 76(4), 564– 575. Kuo, P. X., Carp, J., Light, K. C., & Grewen, K. M. (2012). Neural responses to infants linked with behavioral interactions and testosterone in fathers. Biological Psychiatry, 91(2), 302– 306. doi: 10.1016/j.biopsycho.2012.08.002 Kuzawa, C. W., Gettler, L. T., Muller, M. N., McDade, T. W., & Feranil, A. B. (2009). Fatherhood, pairbonding, and testosterone in the Phillipines. Hormones and Behavior, 56(4), 429–435. Lachance-Grzela, M., & Bouchard, G. (2009). The well-being of cohabiting and married couples during pregnancy: Does pregnancy planning matter? Journal of Social and Personal Relationships, 26(2–3), 141–159. doi: 10.1177/0265407509106705 Lakatos, K., Nemoda, Z., Toth, I., Ronai, Z., Ney, K., Sasvari-Szekely, M., & Gervai, J. (2002). Further evidence for the role of the dopamine D4 receptor (DRD4) gene in attachment disorganization: Interaction of the exon III 48-bp repeat and the-521 C/T promoter polymorphisms. Molecular Psychiatry, 7(1), 27–31. doi: 10.1038/sj/mp/4000986 Lakatos, K., Toth, I., Nemoda, Z., Ney, K., Sasvari-Szekely, M., & Gervai, J. (2000). Dopamine D4 receptor (DRD4) gene polymorphism is associated with attachment disorganization in infants. Molecular Psychiatry, 5(6), 633. doi: 10.1038/sj.mp.4000773

Lambert, K. G., & Kinsley, C. H. (2012). Brain and behavioral modifications that accompany the onset of motherhood. Parenting: Science and Practice, 12(1), 74–89. doi: 10.1080/15295192.2012.638868 Lamborn, S. D., Dornbusch, S. M., & Steinberg, L. (1996). Ethnicity and community context as moderators of the relations between family decision making and adolescent adjustment. Child Development, 67(2), 283–301. Lambrot, R., Xu, C., Saint-Phar, S., Chountal, G., Cohen, T., Paquet, M.,… Kimmins, S. (2013). Low paternal dietary folate alters the mouse sperm epigenome and is associated with negative pregnancy outcomes. Nature Communications, 4, 1–13. doi: 10.1038/ncomms3889 Landry, S. H., Smith, K. E., Swank, P. R., & Guttentag, C. (2008). A responsive parenting intervention: The optimal timing across early childhood for impacting maternal behaviors and child outcomes. Developmental Psychology, 44(5), 1335. doi: 10.1037/a0013030 Langlois, J. H., Ritter, J. M., Casey, R. J., & Sawin, D. B. (1995). Infant attractiveness predicts maternal behaviors and attitudes. Developmental Psychology, 31, 464–472. Lansford, J. E., Alampay, L., Bacchini, D., Bombi, A. S., Bornstein, M. H., Chang, L.,… Uribe Tirado, L. M. (2011). Corporal punishment of children in nine countries as a function of child gender and parent gender. International Journal of Pediatrics, on-line. Lansford, J. E., & Bornstein, M. H. (2007). Review of parenting programs in developing countries. New York: UNICEF, Early Childhood Development Team, August. Lansford, J. E., Bornstein, M. H., Dodge, K. A., Skinner, A., Putnick, D. L., & Deater-Deckard, K. (2011). Attributions and attitudes of mothers and fathers in the United States. Parenting: Science and Practice, 11, 199–213. Lansford, J. E., Chang, L., Dodge, K. A., Malone, P. S., Oburu, P., Palmérus, K.,… Quinn, N. (2005). Physical discipline and children's adjustment: Cultural normativeness as a moderator. Child Development, 76(6), 1234–1246. doi: 10.1111/j.1467–8624.2005.00847.x Lansford, J. E., Criss, M. M., Laird, R. D., Shaw, D. S., Pettit, G. S., Bates, J. E., & Dodge, K. A. (2011). Reciprocal relations between parents' physical discipline and children's externalizing behavior during middle childhood and adolescence. Development and Psychopathology, 23(1), 225–238. doi: 10.1017/S0954579410000751 Lansford, J. E., & Deater-Deckard, K. (2012). Childrearing discipline and violence in developing countries. Child Development, 83(1), 62–75. doi: 10.1111/j.1467– 8624.2011.01676.x Lansford, J. E., Woodlief, D., Malone, P. S., Oburu, P., Pastorelli, C., Skinner, A. T.,… Dodge, K. A. (2014). A longitudinal examination of mothers' and fathers' social information processing biases and harsh discipline in nine countries. Development and Psychopathology, 26, 561– 573.

LaPiere, R. T. (1934). Attitudes versus actions. Social Forces, 13, 230–237. Lareau, A. (2003). Unequal childhoods: Class, race, and family life. Berkeley, CA: University of California Press. Larson, R., & Asmussen, L. (1991). Anger, worry, and hurt in early adolescence: An enlarging world of negative emotions. In M. E. Colten & S. Gore (Eds.), Adolescent stress: Causes and consequences (pp. 21–41). New York, NY: Aldine de Gruyter. Larson, R., Csikszentmihalyi, M., & Graef, R. (1980). Mood variability and the psychosocial adjustment of adolescents. Journal of Youth and Adolescence, 9(6), 469–490. Larson, R., & Lampman-Petraitis, C. (1989). Daily emotional states as reported by children and adolescents. Child Development, 60(5), 1250–1260. Larzelere, R. E. (2000). Child outcomes of nonabusive and customary physical punishment by parents: An updated literature review. Clinical Child and Family Psychology Review, 3(4), 199–221. doi: 10.1023/A:1026473020315 Lau, A. S., Huang, M. M., Garland, A. F., McCabe, K. M., Yeh, M., & Hough, R. L. (2006). Racial variation in self-labeled child abuse and associated internalizing symptoms among adolescents who are high risk. Child Maltreatment, 11(2), 168–181. doi: 10.1177/1077559505285776 Laurent, H. K., & Ablow, J. C. (2012). A cry in the dark: Depressed mothers show reduced neural activation to their own infant's cry. Social Cognitive and Affective Neuroscience, 7(2), 125–34. doi: 10.1093/scan/nsq091 Lawson, D. W., & Mace, R. (2009). Trade-offs in modern parenting: A longitudinal study of sibling competition for parental care. Evolution and Human Development, 30(3), 170–183. doi: 10.1016/j.evolhumbehav.2008.12.001 Leach, P. (2012). Your baby and child: From birth to age five. New York, NY: Knopf. Lecuyer-Maus, E. A. (2000). Maternal sensitivity and responsiveness, limit-setting style, and relationship history in the transition to toddlerhood. Issues in Comprehensive Pediatric Nursing, 23(2), 117–139. doi: 10.1080/01460860050121439 Lee, C., Barr, R. G., Catherine, N., & Wicks, A. (2007). Age-related incidence of publicly reported shaken baby syndrome cases: Is crying a trigger for shaking? Journal of Developmental and Behavioral Pediatrics, 28(4), 288–293. doi: 10.1097/DBP.0b013e3180327b55 Lee, D. M., & Colletta, N. D. (1983, April). Family support for adolescent mothers: The positive and negative impact. Paper presented at the biennial meeting of the Society for Research in Child Development , Detroit, MI. Lee, M. A., Schoppe-Sullivan, S. J., & Kamp Dush, C. M. (2012). Parenting perfectionism and

parental adjustment. Personality and Individual Differences, 52(3), 454–457. doi: 10.1016/j.paid.2011.10.047 Lee, R. (1990). The demographic response to economic crisis in historical and contemporary populations. Population Bulletin of the United Nations, 29, 1–15. Lee, S. S., Chronis-Tuscano, A., Keenan, K., Pelham, W. E., Loney, J., Van Hulle, C. A.,… Lahey, B. B. (2010). Association of maternal dopamine transporter genotype with negative parenting: Evidence for gene × environment interaction with child disruptive behavior. Molecular Psychiatry, 15(5), 548–558. doi: 10.1038/mp.2008.102 Lee, Y. (2009). Early motherhood and harsh parenting: The role of human, social, and cultural capital. Child Abuse & Neglect, 33, 625–637. Leerkes, E. M., Blankson, A. N., & O'Brien, M. (2009). Differential effects of maternal sensitivity to infant distress and nondistress on social-emotional functioning. Child Development, 80(3), 762–775. doi: 10.1111/j.1467–8624.2009.01296.x Leerkes, E. M., & Burney, R. V. (2007). The development of parenting efficacy among new mothers and fathers. Infancy, 12(1), 45–67. doi: 10.1111/j.1532–7078.2007.tb00233.x Leerkes, E. M., & Crockenberg, S. C. (2002). The development of maternal self-efficacy and its impact on maternal behavior. Infancy, 3(2), 227–247. Leerkes, E. M., & Crockenberg, S. C. (2006). Antecedents of mothers' emotional and cognitive responses to infant distress: The role of family, mother, and infant characteristics. Infant Mental Health Journal, 27(4), 405–428. doi: 10.1002/imhj.20099 Leerkes, E. M., Parade, S. H., & Burney, R. V. (2010). Origins of mothers' and fathers' beliefs about infant crying. Journal of Applied Developmental Psychology, 31(6), 467–474. doi: 10.1016/j.appdev.2010.09.003 Leerkes, E., & Siepak, K. (2006). Attachment linked predictors of women's emotional and cognitive responses to infant distress. Attachment & Human Development, 8, 11–32. Leerkes, E. M., Weaver, J. M., & O'Brien, M. (2012). Differentiating maternal sensitivity to infant distress and non-distress. Parenting: Science and Practice, 12(2–3), 175–184. doi: 10.1080/15295192.2012.683353 Lees, D. G., & Ronan, K. R. (2008). Engagement and effectiveness of parent management training (Incredible Years) for solo high-risk mothers: A multiple baseline evaluation. Behaviour Change, 25(02), 109–128. doi: 10.1375/bech.25.2.109 Leibenluft, L., Gobbini, M. I., Harrison, T., & Haxby, J. V. (2004). Mothers' neural activation in response to pictures of their children and other children. Biological Psychiatry, 56(4), 225– 232. doi: 10.1016/j.biopsych.2004.05.017 Leiderman, P. H., Tulkin, S. R., & Rosenfeld, A. (Eds.). (1977). Culture and infancy:

Variations in the human experience. New York, NY: Academic Press. Leinaweaver, J. (2014). Informal kinship-based fostering around the world: Anthropological findings. Child Development Perspectives, 8(3), 131–136. Lengua, L. J. (2006). Growth in temperament and parenting as predictors of adjustment during children's transition to adolescence. Developmental Psychology, 42(5), 819–832. doi: 10.1037/0012–1649.42.5.819 Lengua, L. J., & Kovacs, E. A. (2005). Bidirectional associations between temperament and parenting and the prediction of adjustment problems in middle childhood. Journal of Applied Developmental Psychology, 26(1), 21–38. doi: 10.1016/j.appdev.2004.10.001 Lenzi, D., Trentini, C., Pantano, P., Macaluso, E., Iacoboni, M., Lenzi, G., & Ammaniti, M. (2009). Neural basis of maternal communication and emotional expression processing during infant preverbal stage. Cerebral Cortex, 19(5), 1124–1133. doi: 10.1093/cercor/bhn153 Leon, I. G. (2002). Adoption losses: Naturally occurring or socially constructed? Child Development, 73(2), 652–663. doi: 10.1111/1467–8624.00429 Lerner, J. V., & Galambos, N. L. (1985). Mother role satisfaction, mother–child interaction, and child temperament: A process model. Developmental Psychology, 21(6), 1157. Lerner, J. V., & Galambos, N. L. (1986). Temperament and maternal employment. New Directions for Child Development, 31, 75–88. Letherby, G. (2002). Childless and bereft?: Stereotypes and realities in relation to “voluntary” and “involuntary” childlessness and womanhood. Sociological Inquiry, 72(1), 7–20. doi: 10.1111/1475–682X.00003 Leuner, B., Glasper, E. R., & Gould, E. (2010). Parenting and plasticity. Trends in Neuroscience, 33(10), 465–473. doi: 10.1016/j.tins.2010.07.003 Leventhal, T., & Brooks-Gunn, J. (2005). Neighborhood and gender effects on family processes: Results from the Moving to Opportunity Program. Family Relations, 54(5), 633– 643. doi: 10.1111/j.1741–3729.2005.00347.x Leventhal, T., Dupéré, V., & Shuey, E. (2015). Children in neighborhoods. In M. H. Bornstein & T. Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R. M. Lerner (Editor-in-Chief), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 493–533). Hoboken, NJ: Wiley. Levine, L., García Coll, C. T., & Oh, W. (1985). Determinants of mother-infant interaction in adolescent mothers. Pediatrics, 75(1), 23–29. Levine, M. (2013). Teach your children well: Why values and coping skills matter more than grades, trophies, or “fat envelopes”. New York, NY: HarperCollins.

LeVine, R. A. (2003). Childhood socialization: Comparative studies of parenting, learning and educational change. Hong Kong, China: University of Hong Kong Press. Levitt, M. J., Weber, R. A., & Clark, M. C. (1986). Social network relationships as sources of maternal support and well-being. Developmental Psychology, 22, 310–316 Lewin, A., Mitchell, S. J., & Ronzio, C. R. (2013). Developmental differences in parenting behavior: Comparing adolescent, emerging adult, and adult mothers. Merrill-Palmer Quarterly, 59(1), 23–49. Li, J., Fung, H., Bakeman, R., Rae, K., & Wei, W. (2014). How European American and Taiwanese mothers talk to their children about learning. Child Development, 85(3), 1206– 1221. doi: 10.1111/cdev.12172 Lightfoot, C., & Valsiner, J. (1992). Parental beliefs about developmental processes. Human Development, 25, 192–200. Lino, M. (2013). Expenditures on children by families, 2012. U.S. Department of Agriculture, Center for Nutrition Policy and Promotion. Miscellaneous Publication No. 1528–2012. Livingston, G., & Cohn, D. (2010). Childlessness up among all women; down among women with advanced degrees. Retrieved from Pew Social & Demographic Trends website: http://www.pewsocialtrends.org/2010/06/25/childlessness-up-among-all-women-downamong-women-with-advanced-degrees/ Lloyd, K., & Devine, P. (2006) Parenting practices in Northern Ireland: Evidence from the Northern Ireland Household Panel Survey, Child Care in Practice, 12(4), 365–376. doi: 10.1080/13575270600863275 Locke, J. (1692). Some thoughts concerning education, Part 1 (Harvard Classics, Vol. 37). New York, NY: P. F. Collier & Son. Locke, L. M., & Prinz, R. J. (2002). Measurement of parental discipline and nurturance. Clinical Psychology Review, 22(6), 895–929. doi: 10.1016/S0272–7358(02)00133–2 Loeber, R., Burke, J. D., Lahey, B. B., Winters, A., & Zera, M. (2000). Oppositional defiant and conduct disorder: A review of the past 10 years, Part I. Journal of the American Academy of Child & Adolescent Psychiatry, 39(12), 1468–1484. doi: 10.1097/00004583–200012000– 00007 Lorber, M. F., O'Leary, S. G., & Smith Slep, A. M. (2011). An initial evaluation of the role of emotion and impulsivity in explaining racial/ethnic differences in the use of corporal punishment. Developmental Psychology, 47, 1744–1749. Lorberbaum, J. P., Newman, J. D., Dubno, J. R., Horwitz, A. R., Nahas, Z., Teneback, C. C.,… George, M. S. (1999). Feasibility of using fMRI to study mothers responding to infant cries. Depression and Anxiety, 10(3), 99–104.

Lorenz, K. (1935/1970). Studies in animal and human behavior (R. Martin, Trans.). London, UK: Methuen. Losoya, S. H., Callor, S., Rowe, D. C., & Goldsmith, H. H. (1997). Origins of familial similarity in parenting: A study of twins and adoptive siblings. Developmental Psychology, 33(6), 1012. Lou, V. W., & Chi, I. (2012). Grandparenting roles and functions. In K. K. Mehta & L. L. Thang (Eds.), Experiencing grandparenthood: An Asian perspective (pp. 47–59). Dordrecht, the Netherlands: Springer. doi: 10.1007/978–94–007–2303–0_3 Lounds, J. J., Borkowski, J. G., Whitman, T. L., Maxwell, S. E., & Weed, K. (2005). Adolescent parenting and attachment during infancy and early childhood. Parenting: Science and Practice, 5(1), 91–118. doi: 10.1207/s15327922par0501_4 Lovejoy, M. C., Graczyk, P. A., O'Hare, E., & Neuman, G. (2000). Maternal depression and parenting behavior: A meta-analytic review. Clinical Psychology Review, 20(5), 561–592. doi: 10.1016/S0272–7358(98)00100–7 Ludwig, J., Duncan, G. J., Gennetian, L. A., Katz, L. F., Kessler, R. C., Kling, J. R., & Sanbonmatsu, L. (2012). Neighborhood effects on the long-term well-being of low-income adults. Science, 337(6101), 1505–1510. doi: 10.1126/science.1224648 Lunkenheimer, E. S., Dishion, T. J., Shaw, D. S., Connell, A. M., Gardner, F., Wilson, M. N., & Skuban, E. M. (2008). Collateral benefits of the Family Check-Up on early childhood school readiness: Indirect effects of parents' positive behavior support. Developmental Psychology, 44(6), 1737–1752. doi: 10.1037/a0013858 Luster, T., Rhoades, K., & Haas, B. (1989). The relation between parental values and parenting behavior: A test of the Kohn hypothesis. Journal of Marriage and the Family, 51, 139–147. Luthar, S. S. (2003). The culture of affluence: Psychological costs of material wealth. Child Development, 74(6), 1581–1593. Luthar, S. S., & Barkin, S. H. (2012). Are affluent youth truly “at risk”? Vulnerability and resilience across three diverse samples. Development and Psychopathology, 24(2), 429–449. doi: 10.1017/S0954579412000089 Luthar, S. S., & Latendresse, S. J. (2005). Children of the affluent: Challenges to well-being. Current Directions in Psychological Science, 14(1), 49–53. doi: 10.1111/j.0963– 7214.2005.00333.x Lytton, H., & Gallagher, L. (2002). Parenting twins and the genetics of parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Children and parenting (2nd ed., Vol. 1, pp. 227– 253). Mahwah, NJ: Erlbaum.

Lytton, H., & Romney, D. M. (1991). Parents' differential socialization of boys and girls: A meta analysis. Psychological Bulletin, 109, 267–296. Lytton, H., & Zwimer, W. (1975). Compliance and its controlling stimuli observed in a natural setting. Developmental Psychology, 11, 769–779. Maas, A. J., Vreeswijk, C. M., & van Bakel, H. J. (2013). Effect of situation on mother–infant interaction. Infant Behavior and Development, 36(1), 42–49. doi: 10.1016/j.infbeh.2012.10.006 Maccoby, E. E. (1992). The role of parents in the socialization of children: An historical overview. Developmental Psychology, 28(6), 1006–1017. Maccoby, E. E. (2000). Parenting and its effects on children: On reading and misreading behavior genetics. Annual Review of Psychology, 51(1), 1–27. doi: 10.1146/annurev.psych.51.1.1 Maccoby, E. E., & Martin, J. A. (1983). Socialization in the context of the family: Parent-child interaction. In P. H. Mussen (Series Ed.) & E. M. Hetherington (Vol. Ed.), Handbook of child psychology: Socialization, personality, and social development (4th ed., Vol. 4, pp. 1–101). New York, NY: Wiley. MacDonald, K., & MacDonald, T. M. (2010). The peptide that binds: A systematic review of oxytocin and its prosocial effects in humans. Harvard Review of Psychiatry, 18(1), 1–21. doi: 10.3109/10673220903523615 MacDonald, K., & Parke, R. D. (1984). Bridging the gap: Parent-child play interaction and peer interactive competence. Child Development, 55, 1265–1277. MacKenzie, M. J., Nicklas, E., Brooks-Gunn, J., & Waldfogel, J. (2011). Who spanks infants and toddlers? Evidence from the Fragile Families and Child Well-Being Study. Children and Youth Services Review, 33(8), 1364–1373. doi: 10.1016/j.childyouth.2011.04.007 MacPhee, D. (1981). Manual for the knowledge of infant development inventory. Unpublished manuscript, University of North Carolina, Chapel Hill, NC. Macunovich, D. J. (1996). Relative income and price of time: Exploring their effects on U.S. fertility and female labor force participation. Population Development Review, 22, 223–257. Madge, C., & O'Connor, H. (2006). Parenting gone wired: Empowerment of new mothers on the Internet? Social & Cultural Geography, 7(2), 199–220. doi: 10.1080/14649360600600528 Maestripieri, D. (2005). Early experience affects the intergenerational transmission of infant abuse in rhesus monkeys. Proceedings of the National Academy of Sciences USA, 102(27), 9726–9729. doi: 10.1073/pnas.0504122102 Maestripieri, D., Hoffman, C. L., Anderson, G. M., Carter, C. S., & Higley, J. D. (2009).

Mother-infant interactions in free-ranging rhesus macaques: Relationships between physiological and behavioral variables. Physiology and Behavior, 96(4–5), 613–619. doi: 10.1016/j.physbeh.2008.12.016 Magnuson, K. A., & Duncan, G. J. (2002). Parents in poverty. In M. H. Bornstein (Ed.), Handbook of parenting: Applied parenting (2nd ed., Vol. 4, pp. 95–121). Mahwah, NJ: Erlbaum. Mahoney, A. (2005). Religion and conflict in marital and parent-child relationships. Journal of Social Issues, 61(4), 689–706. doi: 10.1111/j.1540–4560.2005.00427.x Mahoney, A. (2010). Religion in families, 1999–2009: A relational spirituality framework. Journal of Marriage and Family, 72(4), 805–827. doi: 10.1111/j.17413737.2010.00732.x Mahoney, A., Pargament, K., Jewell, T., Swank, A. B., Scott, E., Emery, E., & Rye, M. (1999). Marriage and the spiritual realm: The role of proximal and distal religious constructs in marital functioning. Journal of Family Psychology, 13(3), 321. Mahoney, A., Pargament, K., Swank, A., & Tarakeshwar, N. (2001). Religion in the home in the 1980's and 90's: A review and conceptual integration of empirical links between religion, marriage, and parenting. Journal of Family Psychology, 15, 559–596. Mahoney, A., Pargament, K., Tarakeshwar, N., & Swank, A. B. (2008). Religion in the home in the 1980s and 1990s: A meta-analytic review and conceptual analysis of links between religion, marriage, and parenting. Psychology of Religion and Spirituality, S(1), 63–101. doi: 10.1037/1941–1022.S.1.63 Mammen, O. K., Kolko, D. J., & Pilkonis, P. A. (2002). Negative affect and parental aggression in child physical abuse. Child Abuse & Neglect, 26, 407–424. doi: 10.1016/S0145-2134(02)00316-2 Manian, N., & Bornstein, M. H. (2009). Dynamics of emotion in infants of clinically depressed and nondepressed mothers. Journal of Child Psychology and Psychiatry, 50(11), 1410–1418. doi: 10.1111/j.1469–7610.2009.02166.x Mansbach, I. K., & Greenbaum, C. W. (1999). Developmental maturity expectations of Israeli fathers and mothers: Effects of education, ethnic origin, and religiosity. International Journal of Behavioral Development, 23, 771–797. Mantzicopoulos, P. Y. (1997). The relationship of family variables to Head Start children's preacademic competence. Early Education and Development, 8, 357–375. Markus, H., Cross, S., & Wurf, E. (1990). The role of the self-system in competence. In R. J. Sternberg & J. Kolligian, Jr. (Eds.), Competence considered (pp. 205–225). New Haven, CT: Yale University Press. Marlowe, F. W. (2003). The mating system of foragers in the standard cross-cultural sample.

Cross-Cultural Research, 37(3), 282–306. doi: 10.1177/1069397103254008 Marquardt, E., Blankenhorn, D., Lerman, R. I., Malone-Colón, L., & Bradford Wilcox, W. (2012). The President's marriage agenda for the forgotten sixty percent: The state of our unions. Charlottesville, VA: National Marriage Project and Institute for American Values. Martin, A., Ryan, R. M., & Brooks-Gunn, J. (2007). The joint influence of mother and father parenting on child cognitive outcomes at age 5. Early Childhood Research Quarterly, 22(4), 423–439. Martin, J. A., Hamilton, B. E., Ventura, S. J., Osterman, M. J. K., & Matthews, T. J. (2013). Births: Final data for 2011. National Vital Statistics Reports (Vol. 62). Hyattsville, MD: National Center for Health Statistics. Martin, S. E., Clements, M. L., & Crnic, K. A. (2002). Maternal emotions during mothertoddler interaction: Parenting in affective context. Parenting: Science and Practice, 2(2), 105–126. doi: 10.1207/S15327922PAR0202_02 Mascaro, J. S., Hackett, P. D., & Rilling, J. K. (2013). Testicular volume is inversely correlated with nurturing-related brain activity in human fathers. Proceedings of the National Academy of Sciences, 110(39), 15746–15751. Mattson, B. J., Williams, S., Rosenblatt, J. S., & Morrell, J. I. (2001). Comparison of two positive reinforcing stimuli: Pups and cocaine throughout the postpartum period. Behavioral Neuroscience, 115(3), 683–694. doi: 10.1037/0735–7044.115.3.683 McAdams, D. P., & Pals, J. L. (2006). A new Big Five: Fundamental principles for an integrative science of personality. American Psychologist, 61(3), 204–217. doi: 10.1037/0003–066X.61.3.204 McCall, R. B., & Green, B. L. (2004). Beyond the methodological gold standards of behavioral research: Considerations for practice and policy. Society for Research in Child Development Social Policy Reports, 18, 1–19. McCartney, K., Harris, M. J., & Bernieri, F. (1990). Growing up and growing apart: A developmental meta-analysis of twin studies. Psychological Bulletin, 107(2), 226–237. McConnell, D., Breitkreuz, R., & Savage, A. (2011). From financial hardship to child difficulties: Main and moderating effects of perceived social support. Child: Care, Health and Development, 37(5), 679–691. doi: 10.1111/j.1365–2214.2010.01185.x McCrae, R. R., Terraciano, A., & 78 members of the Personality Profiles of Cultures Project. (2005). Universal features of personality traits from the observer's perspective: Data from 50 cultures. Journal of Personality and Social Psychology, 88(3), 547–561. doi: 10.1037/0022– 3514.88.3.547 McGillicuddy-DeLisi, A. V. (1992). Parents' beliefs and children's personal-social

development. In I. Sigel & A. V. McGillicuddy-DeLisi (Eds.), Parental belief systems: The psychological consequences for children (2nd ed., pp. 115–142). Hillsdale, NJ: Erlbaum. McGrath, J. J., Petersen, L., Agerbo, E., Mors, O., Mortensen, P. B., & Pedersen, C. B. (2014). A comprehensive assessment of parental age and psychiatric disorders. JAMA Psychiatry, 71, 301–309. McGue, M., Elkins, I., Walden, B., & Iacono, W. G. (2005). Perceptions of the parentadolescent relationship: A longitudinal investigation. Developmental Psychology, 41(6), 971. doi: 10.1037/0012–1649.41.6.971 McGuire, S. (2003). The heritability of parenting. Parenting: Science and Practice, 3(1), 73– 94. doi: 10.1207/S15327922PAR0301_04 McGuire, S., Segal, N. L., & Hershberger, S. (2012). Parenting as phenotype: A behavioral genetic approach to understanding parenting. Parenting: Science and Practice, 12(2–3), 192– 201. doi: 10.1080/15295192.2012.683357 McHale, J., Khazan, I., Erera, P., Rotman, T., DeCourcey, W., & McConnell, M. (2002). Coparenting in diverse family systems. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 75–107). Mahwah, NJ: Erlbaum. McHale, J., & Lindahl, K. (2011). Coparenting: A conceptual and clinical examination of family systems. Washington, DC: American Psychological Association Press. doi: 10.1037/12328–000 McHale, S. M., & Pawletko, T. M. (1992). Differential treatment of siblings in two family contexts. Child Development, 63(1), 68–81. McIntosh, J., Lyon, R., Carlson, G., Everette, C., & Loera, S. (2008). Measuring the mesosystem: A survey and critique of approaches to cross setting measurement for ecological research and models of collaborative care. Family Systems and Health, 26, 86–104. McLaughlin, B., White, D., McDevitt, T., & Raskin, R. (1983). Mothers' and fathers' speech to their young children: Similar or different? Journal of Child Language, 10, 245–252. McLoyd, V. C., Aikens, N. L., & Burton, L. M. (2006). Childhood poverty, policy, and practice. In K. A. Renninger & I. E. Sigel (Vol. Eds.), Child psychology in practice. In W. Damon & R. M. Lerner (Editors-in-Chief), Handbook of child psychology (6th ed., Vol. 4, pp. 700–775). Hoboken, NJ: Wiley. doi: 10.1002/9780470147658.chpsy0418 Mejia, R., Miewer, W., & Williams, L. (2006). Domestic violence exposure in Colombianadolescents: Pathways to violent and prosocial behavior. Journal of Traumatic Stress, 19(2), 257–267. Mekos, D., Hetherington, E. M., & Reiss, D. (1996). Sibling differences in problem behavior

and parental treatment in nondivorced and remarried families. Child Development, 67(5), 2148–2165. Melamed, B. G. (2002). Parenting the ill child. In M. H. Bornstein (Ed.), Handbook of parenting (Vol. 5, pp. 329–348). Mahwah, NJ: Erlbaum. Mendez, J. L., Carpenter, J. L., LaForett, D. R., & Cohen, J. S. (2009). Parental engagement and barriers to participation in a community-based preventive intervention. American Journal of Community Psychology, 44(1–2), 1–14. doi: 10.1007/s10464–009–9252-x Mendl, M. (1988). The effects of litter-size variation on the development of play behaviour in the domestic cat: Litters of one and two. Animal Behaviour, 36(1), 20–34. Mertesacker, B., Bade, U., Haverkock, A., & Pauli-Pott, U. (2004). Predicting maternalreactivity/sensitivity: The role of infant emotionality, maternal depressiveness/anxiety, and social support. Infant Mental Health Journal, 25(1), 47–61. doi: 10.1002/imhj.10085 Mesman, J., Van IJzendoorn, M. H., & Bakermans-Kranenburg, M. J. (2012). Unequal in opportunity, equal in process: Parental sensitivity promotes positive child development in ethnic minority families. Child Development Perspectives, 6, 239–250. doi: 10.1111/j.17508606.2011.00223.x Mesman, J., Van IJzendoorn, M. H., Behrens, K., Carbonell, O. A., Cárcamo, R. A., CohenParaira, I.,… Zreik, G. (2015). Is the ideal mother a sensitive mother? Beliefs about early childhood parenting in mothers across the globe. International Journal of Behavioral Development. Meteyer, K., & Perry-Jenkins, M. (2010). Father involvement among working-class, dualearner couples. Fathering: A Journal of Theory, Research, and Practice about Men as Fathers, 8(3), 379–403. doi: 10.3149/fth.0803.379 Metsäpelto, R. L., & Pulkkinen, L. (2003). Personality traits and parenting: Neuroticism, extraversion, and openness to experience as discriminative factors. European Journal of Personality, 17(1), 59 –78. doi: 10.1002/per.468 Meunier, J. C., Roskam, I., & Browne, D. T. (2011). Relations between parenting and child behavior: Exploring the child's personality and parental self-efficacy as third variables. International Journal of Behavioral Development, 35(3), 246–259. doi: 10.1177/0165025410382950 Mikulincer, M., & Shaver, P. R. (2010). Attachment in adulthood: Structure, dynamics, and change. New York, NY: Guilford Press. Mikulincer, M., Shaver, P. R., & Pereg, D. (2003). Attachment theory and affect regulation: The dynamics, development, and cognitive consequences of attachment-related strategies. Motivation and Emotion, 27(2), 77–102. doi: 10.1023/A:1024515519160

Mileva-Seitz, V., Kennedy, J., Atkinson, L., Steiner, M., Levitan, R., Matthews, S. G.,… Fleming, A. S. (2011). Serotonin transporter allelic variation in mothers predicts maternal sensitivity, behavior, and attitudes toward 6-month-old infants. Genes, Brain, and Behavior, 10(3), 325–333. doi: 10.1111/j.1601–183X.2010.00671.x Miller, A. L., McDonough, S. C., Rosenblum, K. L., & Sameroff, A. J. (2002). Emotion regulation in context: Situational effects on infant and caregiver behavior. Infancy, 3, 403–433. Miller, P. J., Wang, S. H., Sandel, T., & Cho, G. E. (2002). Self-esteem as folk theory: A comparison of European American and Taiwanese mother's beliefs. Parenting: Science and Practice, 2(3), 209–239. doi: 10.1207/S15327922PAR0203_02 Miller, S. G. (1988). Parents' beliefs about children's cognitive development. Child Development, 59, 259–285. Mills-Koonce, W. G. J., Propper, C., & Al-Yagon, C. (2007). Infant and parent factors associated with early maternal sensitivity: A caregiver-attachment systems approach. Infant Behavior & Development, 30(1), 114–126. doi: 10.1016/j.infbeh.2006.11.010 Minuchin, P. (1985). Families and individual development: Provocations from the field of family therapy. Child Development, 56, 289–302. Mirowsky, J. (2002). Parenthood and health: The pivotal and optimal age at first birth. Social Forces, 81(1), 315–349. doi: 10.1353/sof.2002.0055 Moller, A. P. (2003). The evolution of monogamy: Mating relationships, parental care, and sexual selection. In C. Boesch (Ed.), Monogamy: Mating strategies and partnerships in birds, humans, and other mammals (pp. 29–41). Cambridge, UK: Cambridge University Press. Molnar, B. E., Buka, S. L., Brennan, R. T., Holton, J. K., & Earls, F. (2003). A multilevel study of neighborhoods and parent-to-child physical aggression: Results from the project on human development in Chicago neighborhoods. Child Maltreatment, 8(2), 84–97. doi: 10.1177/1077559502250822 Mondschein, E. R., Adolph, K. E., & Tamis-LeMonda, C. S. (2000). Gender bias in mothers' expectations about infant crawling. Journal of Experimental Child Psychology, 77(4), 304– 316. doi: 10.1006/jecp.2000.2597 Moore, M. R., & Brooks-Gunn, J. (2002). Adolescent parenthood. In M. H. Bornstein (Ed.),Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 173–214). Mahwah, NJ: Erlbaum. Moran, G., Pederson, D. R., Pettit, P., & Krupka, A. (1992). Maternal sensitivity and infant mother attachment in a developmentally delayed sample. Infant Behavior and Development, 15(4), 427–442.

Morelli, G. A., & Rothbaum, F. (2007). Situating the child in context. Attachment Relationships and Self-Regulation in Different Cultures, 20, 500–527. Mrug, S., & Windle, M. (2009). Mediators of neighborhood influences on externalizing behavior in preadolescent children. Journal of Abnormal Child Psychology, 37(2), 265–280. doi: 10.1007/s10802–008–9274–0 Mulvaney, M. K., McCartney, K., Bub, K. L., & Marshall, N. L. (2006). Determinants of dyadic scaffolding and cognitive outcomes in first graders. Parenting: Science and Practice, 6(4), 297–320. doi: 10.1207/s15327922par0604_2 Murdock, G. P. (1959). Cross-language parallels in parental kin terms. Anthropological Linguistics, 1(9), 1–5. Murkoff, H., & Mazel, S. (2008). What to expect when you're expecting. New York, NY: Workman. Murray, L., Halligan, S. L., & Cooper, P. J. (2010). Effects of postnatal depression on mother infant interactions and child development. In J. G. Bremner & T. D. Wachs (Eds.), Handbook of infant development: Applied and policy issues (Vol. 2, pp. 192–220). Malden, MA: WileyBlackwell. doi: 10.1002/9781444327588.ch8 Murray-Swank, A., Mahoney, A., & Pargament, K. I. (2006). Sanctification of parenting: Links to corporal punishment and parental warmth among Biblically conservative and liberal mothers. International Journal for the Psychology of Religion, 16, 271–288. doi: 10.1207/s15327582ijpr1604_3 Mytton, J., Ingram, J., Manns, S., & Thomas, J. (2014). Facilitators and barriers to engagement in parenting programs: A qualitative systematic review. Health Education & Behavior, 41(2), 127–137. doi: 10.1177/1090198113485755 Na, J. C., & Chia, S. W. (2008). Impact of online resources on informal learners: Parents' perception of their parenting skills. Computers & Education, 51(1), 173–186. doi: 10.1016/j.compedu.2007.05.006 Naber, F., van IJzendoorn, M. H., Deschamps, P., van Engeland, H., & Bakermans-Kranenburg, M. J. (2010). Intranasal oxytocin increases fathers' observed responsiveness during play with their children: A double-blind within-subject experiment. Psychoneuroendocrinology, 35(10), 1583–1586. doi: 10.1016/j.psyneuen.2010.04.007 National Research Council. (2013). New directions in child abuse and neglect research. Retrieved from http://www.iom.edu/Reports/2013/New-Directions-in-Child-Abuse-andNeglect-Research.aspx? utm_source=feedburner&utm_medium=feed&utm_campaign=Feed%3A+NewIomReports+ (New+IOM+Reports) Natsuaki, M. N., Leve, L. D., Harold, G. T., Neiderhiser, J. M., Shaw, D. S., Ganiban, J.,…

Reiss, D. (2013). Transactions between child social wariness and observed structured parenting: Evidence from a prospective adoption study. Child Development, 84(5), 1750– 1765. doi: 10.1111/cdev.12070 Naumova, O. Y., Lee, M., Rychkov, S. Y., Vlasova, N. V., & Grigorenko, E. L. (2013). Gene expression in the human brain: The current state of the study of specificity and spatiotemporal dynamics. Child Development, 84(1), 76–88. doi: 10.1111/cdev.12014 Neiderhiser, J. M., Reiss, D., Pedersen, N. L., Lichtenstein, P., Spotts, E. L., Hansson, K.,… Elthammer, O. (2004). Genetic and environmental influences on mothering of adolescents: A comparison of two samples. Developmental Psychology, 40(3), 335. doi: 10.1037/0012– 1649.40.3.335 Neitzel, C., & Stright, A. D. (2004). Parenting behaviours during child problem solving: The roles of child temperament, mother education and personality, and the problem-solving context. International Journal of Behavioral Development, 28(2), 166–179. doi: 10.1080/01650250344000370 Nelson, S. K., Kushlev, K., English, T., Dunn, E. W., & Lyubomirsky, S. (2013). In defense of parenthood children are associated with more joy than misery. Psychological Science, 24(1), 3–10. doi: 10.1177/0956797612447798 Neppl, T. K., Conger, R. D., Scaramella, L. V., & Ontai, L. L. (2009). Intergenerational continuity in parenting behavior: Mediating pathways and child effects. Developmental Psychology, 45, 1241–1256. doi: 10.1037/a0014850 Newcomb, M. D., & Locke, T. F. (2001). Intergenerational cycle of maltreatment: A popular concept obscured by methodological limitations. Child Abuse and Neglect, 25(9), 1219–1240. doi: 10.1016/S0145–2134(01)00267–8 Ng, F. F. Y., Tamis-LeMonda, C. S., Godfrey, E. B., Hunter, C. J., & Yoshikawa, H. (2012). Dynamics of mothers' goals for children in ethnically diverse populations across the first three years of life. Social Development, 21(4), 821–848. doi: 10.1111/j.1467–9507.2012.00664.x Niccols, A., & Feldman, M. (2006). Maternal sensitivity and behaviour problems in young children with developmental delay. Infant and Child Development, 15(5), 543–554. doi: 10.1002/icd.468 Nichols, S., Nixon, H., Pudney, V., & Jurvansuu, S. (2009). Parents resourcing children's early development and learning. Early Years, 29(2), 147–161. doi: 10.1080/09575140902831387 Ninio, A. (1980). Picture-book reading in mother-infant dyads belonging to two subgroups in Israel. Child Development, 51, 587–590. Nitschke, J. B., Nelson, E. E., Rusch, B. D., Fox, A. S., Oakes, T. R., & Davidson, R. J. (2004). Orbitofrontal cortex tracks positive mood in mothers viewing pictures of their newborn infants. Neuroimage, 21(2), 583–592. doi: 10.1016/j.neuroimage.2003.10.005

Noriuchi, M., Kikuchi, Y., & Senoo, A. (2008). The functional neuroanatomy of maternal love: Mother's response to infant's attachment behaviors. Biological Psychiatry, 63(4), 415–423. doi: 10.1016/j.biopsych.2007.05.018 Numan, M. (2012). Maternal behavior: Neural circuits, stimulus valence, and motivational processes. Parenting: Science and Practice, 12(2–3), 105–114. doi: 10.1080/15295192.2012.680406 Numan, M., Fleming, A. S., & Levy, F. (2006). Maternal behavior. In J. D. Neill (Ed.), Knobil and Neill's physiology of reproduction (3rd ed., Vol. 2, pp. 1921–1993). Boston, MA: Elsevier. doi: 10.1016/B978–012515400–0/50040–3 Numan, M., & Insel, T. R. (2003). The neurobiology of parental behavior (Vol. 1). New York, NY: Springer. Numan, M., & Stolzenberg, D. S. (2009). Medial preoptic area interactions with dopamine neural systems in the control of the onset and maintenance of maternal behavior in rats. Frontiers in Neuroendocrinology, 30(1), 46–64. doi: 10.1016/j.yfrne.2008.10.002 O'Brien Caughy, M., & Franzini, L. (2005). Neighborhood correlates of cultural differences in perceived effectiveness of parental disciplinary actions. Parenting: Science and Practice, 5(2), 119–151. doi: 10.1207/s15327922par0502_1 O'Connell, M. E., Boat, T., & Warner, K. E. (2009). Preventing mental, emotional, and behavioral disorders among young people: Progress and possibilities. Washington, DC: Institute of Medicine, National Research Council. O'Connor, T. G. (2002). The effects of parenting reconsidered: Findings, challenges, and applications. Journal of Child Psychology and Psychiatry, 43(5), 555–572. doi: 10.1111/1469–7610.00046 O'Connor, T. G., Deater-Deckard, K., Fulker, D., Rutter, M., & Plomin, R. (1998). Genotype environment correlations in late childhood and early adolescence: Antisocial behavioral problems and coercive parenting. Developmental Psychology, 34(5), 970–981. O'Connor, T. G., Dunn, J., Jenkins, J. M., & Rasbash, J. R. (2006). Predictors of between family and within-family variation in parent-child relationships. Journal of Child Psychology and Psychiatry, 47, 498–510. Ochs, E. (1988). Culture and language development: Language acquisition and language socialization in a Samoan village. Cambridge, UK: Cambridge University Press. Ogbu, J. U. (1981). Origins of human competence: A cultural-ecological perspective. Child Development, 52, 413–429. Okagaki, L., & Bingham, G. E. (2005). Parents' social cognitions and their parenting behaviors. In T. Luster & L. Okagaki (Eds.), Parenting: An ecological perspective (2nd ed.,

pp. 3–33). Mahwah, NJ: Erlbaum. Olds, D. L., Robinson, J., O'Brien, R., Luckey, D. W., Pettitt, L. M., Henderson, C. R.,… Talmi, A. (2002). Home visiting by paraprofessionals and by nurses: a randomized, controlled trial. Pediatrics, 110(3), 486–496. doi: 10.1542/peds.110.3.486 Östberg, M., & Hagekull, B. (2000). A structural modeling approach to the understanding of parenting stress. Journal of Clinical Child Psychology, 29(4), 615–625. doi: 10.1207/S15374424JCCP2904_13 Östberg, M., Hagekull, B., & Hagelin, E. (2007). Stability and prediction of parenting stress. Infant and Child Development, 16(2), 207–223. doi: 10.1002/icd.516 Pajulo, M., Suchman, N., Kalland, M., Sinkkonen, J., Helenius, H., & Mayes, L. C. (2009). Role of the maternal reflective ability for substance abusing mothers. Journal of Prenatal & Perinatal Psychology & Health, 23, 13–31. Palacios, J., & Moreno, M. C. (1996). Parents' and adolescents' ideas on children: Origins and transmission of intracultural diversity. In S. Harkness & C. M. Super (Eds.), Parents' cultural belief systems: Their origins, expressions, and consequences (pp. 215–253). New York, NY: Guilford Press. Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions. New York, NY: Oxford University Press. Papoušek, H., & Bornstein, M. H. (1992). Didactic interactions: Intuitive parental support of vocal and verbal development in human infants. In H. Papoušek, U. Jurgens, & M. Papoušek (Eds.), Nonverbal vocal communication: Comparative and developmental approaches— Studies in emotion and social interaction (pp. 209–229). New York, NY: Cambridge University Press. Papoušek, H., & Papoušek, M. (2002). Intuitive parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Biology and ecology of parenting (2nd ed., Vol. 2, pp. 183–203). Mahwah, NJ: Erlbaum. Pargament, K. I. (1997). The psychology of religion and coping: Theory, practice and research. New York, NY: Guilford Press. Pargament, K. I., Smith, B. W., Koenig, H. G., & Perez, L. (1998). Patterns of positive and negative religious coping with major life stressors. Journal for the Scientific Study of Religion, 37(4), 710–724. Park, H., Coello, J. A., & Lau, A. S. (2014). Child socialization goals in East Asian versus Western nations from 1989 to 2010: Evidence for social change in parenting. Parenting: Science and Practice, 14(2), 69–91. doi: 10.1080/15295192.2014.914345 Park, H., Twenge, J. M., & Greenfield, P. M. (2013). The Great Recession: Implications for

adolescent values and behavior. Social Psychological and Personality Science, 5(3), 310– 318. doi: 10.1177/1948550613495419 Parke, R. D. (2002). Fathers and families. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 27–73). Mahwah, NJ: Erlbaum. Parke, R., & Buriel, R. (2006). Socialization in the family: Ethnic and ecological perspectives. In N. Eisenberg (Ed.), The handbook of child psychology: Social, emotional, and personality development (6th ed., Vol. 3, pp. 429–504). New York, NY: Wiley. doi: 10.1002/9780470147658.chpsy0308 Parker, G. (1983). Parental overprotection. A risk factor in psychosocial development. New York, NY: Grune and Stratton. Parks, P. L., & Smeriglio, V. L. (1986). Relationships among parenting knowledge, quality of stimulation in the home and infant development. Family Relations: Journal of Applied Family and Child Studies, 35, 411–416. Parsons, C. E., Young, K. S., Kumari, N., Stein, A., & Kringelbach, M. L. (2011). The motivational salience of infant faces is similar for men and women. PLoS One, 6(5), e20632. doi: 10.1371/journal.pone.0020632 Pasley, K., & Gecas, V. (1984). Stresses and satisfactions of the parental role. Personnel and Guidance Journal, 62(7), 400–404. Patterson, G. R., & Forgatch, M. S. (1995). Predicting future clinical adjustment from treatment outcome and process variables. Psychological Assessment, 7(3), 275–285. Paulson, J. F., Bazemore, S. D., & Sharnail, D. (2010). Prenatal and postpartum depression in fathers and its association with maternal depression: A meta-analysis. Journal of the American Medical Association, 303(19), 1961–1969. doi: 10.1001/jama.2010.605 Paulussen-Hoogeboom, M. C., Stams, G. J., Hermanns, J. M. A., & Peetsma, T. D. (2007). Child negative emotionality and parenting from infancy to preschool: A meta-analytic review. Developmental Psychology, 43(2), 438–453. doi: 10.1037/0012–1649.43.2.438 Pears, K. C., & Capaldi, D. M. (2001). Intergenerational transmission of abuse: A two generational prospective study of an at-risk sample. Child Abuse & Neglect, 25(11), 1439– 1461. doi: 10.1016/S0145–2134(01)00286–1 Pearson, R. M., Lightman, S. L., & Evans, J. (2011). Attentional processing of infant emotion during late pregnancy and mother-infant relations after birth. Archives of Women's Mental Health, 14(1), 23–31. doi: 10.1007/s00737–010–0180–4 Pearson, R. M., Melotti, R., Heron, J., Joinson, C., Stein, A., Ramchandani, P. G., & Evans, J. (2012). Disruption to the development of maternal responsiveness? The impact of prenatal depression on mother-infant interactions. Infant Behavior and Development, 35(4), 613–626.

doi: 10.1016/j.infbeh.2012.07.020 Peltonen, K., Ellonen, N., Larsen, H. B., & Helweg-Larsen, K. (2010). Parental violence and adolescent mental health. European Child & Adolescent Psychiatry, 19(11), 813–822. doi: 10.1007/s00787–010–0130–8 Pener-Tessler, R., Avinun, R., Uzefovsky, F., Edelman, S., Ebstein, R. P., & Knafo, A. (2013). Boys' serotonin transporter genotype affects maternal behavior through self-control: A case of evocative gene–environment correlation. Development and Psychopathology, 25(1), 151– 162. doi: 10.1017/S095457941200096X Pennington, B. F., Bennetto, L., McAleer, O., & Roberts, R. J., Jr. (1996). Executive functions and working memory: Theoretical and measurement issues. In R. Lyon & N. A. Krasnegor (Eds.), Attention, memory, and executive function (pp. 327–348). Baltimore, MD: Paul H. Brookes. Pereira, M., Seip, K. M., & Morrell, J. I. (2008). Maternal motivation and its neural substrate across the postpartum period. In R. S. Bridges (Ed.), Neurobiology of the parental brain. New York, NY: Academic Press. Perrin, E. C., Siegel, B. S., & the Committee on Psychosocial Aspects of Child and Family Health (2013). Promoting the well-being of children whose parents are gay or lesbian. Pediatrics, 131(4), 827–830. doi: 10.1542/peds.2013–0376 Perry-Jenkins, M., Repetti, R. L., & Crouter, A. C. (2000). Work and family in the 1990s. Journal of Marriage and the Family, 62(4), 981–998. doi: 10.1111/j.1741– 3737.2000.00981.x Pianta, R., Sroufe, L. A., & Egeland, B. (1989). Continuity and discontinuity in maternal sensitivity at 6, 24 and 42 months in a high risk sample. Child Development, 60(2), 481–487. Pinderhughes, E. E., Dodge, K. A., Bates, J. E., Pettit, G. S., & Zelli, A. (2000). Discipline responses: Influences of parents' socioeconomic status, ethnicity, beliefs about parenting, stress, and cognitive-emotional processes. Journal of Family Psychology, 14(3), 380–400. doi: 10.1037/0893–3200.14.3.380 Pinquart, M., & Teubert, D. (2010). Effects of parenting education with expectant and new parents: A meta-analysis. Journal of Family Psychology, 24(3), 316–327. doi: 10.1037/a0019691 Plomin, R. (2012). Child development and molecular genetics: 14 years later. Child Development, 84(1), 104–120. doi: 10.1111/j.1467–8624.2012.01757.x Plomin, R., DeFries, J. C., McClearn, G. E., & McGuffin, P. (2008). Behavioral genetics (5th ed.). New York, NY: Worth. Plomin, R., Reiss, D., Hetherington, E. M., & Howe, G. W. (1994). Nature and nurture:

Genetic contributions to measures of the family environment. Developmental Psychology, 30(1), 32. Poehlmann, J., Park, J., Bouffiou, L., Abrahams, J., Shlafer, R., & Hahn, E. (2008). Representations of family relationships in children living with custodial grandparents. Attachment & Human Development, 10(2), 165–188. doi: 10.1080/14616730802113695 Pomerantz, E. M., Wang, Q., & Ng, F. F. Y. (2005). Mothers' affect in the homework context: The importance of staying positive. Developmental Psychology, 41(2), 414–427. doi: 10.1037/0012–1649.41.2.414 Pomerleau, A., Bolduc, D., Malcuit, G., & Cossette, L. (1990). Pink or blue: Environmental gender stereotypes in the first two years of life. Sex Roles, 22(5–6), 359–367. Pomerleau, A., Scuccimarri, C., & Malcuit, G. (2003). Mother-infant behavioral interactions in teenage and adult mothers during the first 6 months postpartum: Relations with infant development. Infant Mental Health Journal, 24(5), 495–509. doi: 10.1002/imhj.10073 Porter, C. L., & Hsu, H. C. (2003). First-time mothers' perceptions of efficacy during the transition to motherhood: Links to infant temperament. Journal of Family Psychology,17(1), 54–64. doi: 10.1037/0893–3200.17.1.54 Posada, G., Gao, Y., Wu, F., Posada, R., Tascon, M., Schöelmerich, A.,… Synnevaag, B. (1995). The secure-base phenomenon across cultures: Children's behavior, mother's preferences, and experts' concepts. Monographs of the Society for Research in Child Development, 60(2–3), 27–48. doi: 10.1111/j.1540-5834.1995.tb00202.x Powell, D. R. (2013). Parenting intervention outcome studies: Research design considerations. Parenting: Science and Practice, 13(4), 266–284. doi: 10.1080/15295192.2013.832571 Power, T. G. (2004). Stress and coping in childhood: The parents' role. Parenting: Science and Practice, 4(4), 275–321. doi: 10.1207/s15327922par0404_1 Power, T. G., & Hill, L. G. (2008). Maternal protectiveness and child adjustment: A multidimensional study. Parenting: Science and Practice, 8(3), 187–212. doi: 10.1080/15295190802045428 Price, T. S., & Jaffee, S. R. (2008). Effects of the family environment: Gene-environment interaction and passive gene-environment correlation. Developmental Psychology, 44(2), 305. doi: 10.1037/0012–1649.44.2.305 Prinz, R. J., Sanders, M. R., Shapiro, C. J., Whitaker, D. J., & Lutzker, J. R. (2009). Population based prevention of child maltreatment: The US Triple P system population trial. Prevention Science, 10(1), 1–12. doi: 10.1007/s11121–009–0123–3 Prinzie, P., Stams, G. J. J. M., Deković, M., Reijntjes, A. H. A., & Belsky, J. (2009). The relations between parents' Big Five personality factors and parenting: A meta-analytic review.

Journal of Personality and Social Psychology, 97(2), 351–362. doi: 10.1037/a0015823 Porter, C. L., & Hsu, H. C. (2003). First-time mothers' perceptions of efficacy during the transition to motherhood: Links to infant temperament. Journal of Family Psychology, 17(1), 54–64. doi: 10.1037/0893–3200.17.1.54 Purhonen, M., Valkonen-Korhonen, M., & Lehtonen, J. (2008) The impact of stimulus type and early motherhood on attentional processing. Developmental Psychobiology, 50(6), 600–607. Putallaz, M., Costanzo, P. R., & Smith, R. B. (1991). Maternal recollections of childhood peer relationships: Implications for their children's social competence. Journal of Social and Personal Relationships, 8(3), 403–422. Putnam, S. P., Sanson, A. V., & Rothbart, M. K. (2002). Child temperament and parenting. In M. H. Bornstein (Ed.), Handbook of parenting: Children and parenting (2nd ed., Vol. 1, pp. 255–277). Mahwah, NJ: Erlbaum. Putnick, D. L., Bornstein, M. H., Hendricks, C., Painter, K. M., Suwalsky, J. T. D., & Collins, W. A. (2008). Parenting stress, perceived parenting behaviors, and adolescent self-concept in European American families. Journal of Family Psychology, 22(5), 752–762. doi: 10.1037/a0013177 Putnick, D. L., Bornstein, M. H., Hendricks, C., Painter, K. M., Suwalsky, J. T. D., & Collins, W. A. (2010). Stability, continuity, and similarity of parenting stress in European American mothers and fathers across their child's transition to adolescence. Parenting: Science and Practice, 10(1), 60–77. doi: 10.1080/15295190903014638 Putnick, D. L., Bornstein, M. H., Lansford, J. E., Chang, L., Deater-Deckard, K., Di Giunta, L., … Bombi, A. S. (2012). Agreement in mother and father acceptance-rejection, warmth, and hostility/rejection/neglect of children across nine countries. Cross-Cultural Research: The Journal of Comparative Social Science, 46, 191–223. Quinlan, R. J. (2007). Human parental effort and environmental risk. Proceedings of the Royal Society B: Biological Sciences, 274(1606), 121–125. doi: 10.1098/rspb.2006.3690 Quinn, N. (2005). Universals of child rearing. Anthopological Theory, 5, 477–516. doi: 10.1177/1463499605059233 Rabain-Jamin, J. (1994). Language and socialization of the child in African families living in France. In P. M. Greenfield & R. Cocking (Eds.), Cross-cultural roots of minority child development (pp. 147–166). Hillsdale, NJ: Erlbaum. Racz, S. J., McMahon, R. J., & Luthar, S. S. (2011). Risky behavior in affluent youth: Examining the co-occurrence and consequences of multiple problem behaviors. Journal of Child and Family Studies, 20(1), 120–128. doi: 10.1007/s10826–010–9385–4 Radesky, J. S., Kistin, C. J., Zuckerman, B., Nitzberg, K., Gross, J., Kaplan-Sanoff, M.,…

Silverstein, M. (2014). Patterns of mobile device use by caregivers and children during meals in fast food restaurants. Pediatrics, 133, e843–e849. doi: 10.1542/peds.2013–3703 Radey, M., & Randolph, K. A. (2009). Parenting sources: How do parents differ in their efforts to learn about parenting? Family Relations, 58(5), 536–548. doi: 10.1111/j.1741– 3729.2009.00573.x Ragozin, A. S., Basham, R. B., Crnic, K. A., Greenberg, M. T., & Robinson, N. M. (1982). Effects of maternal age on parenting role. Developmental Psychology, 18(4), 627. Raikes, H., Pan, L., Luze, G., Tamis-LeMonda, C. S., Brooks-Gunn, J., Constantine J.,… Rodriguez, E. T. (2006). Mother-child bookreading in low-income families: Correlates and outcomes during the first three years of life. Child Development, 77(4), 924–953. doi: 10.1111/j.1467–8624.2006.00911.x Randolph, K. A., Fincham, F., & Radey, M. (2009). A framework for engaging parents in prevention. Journal of Family Social Work, 12(1), 56–72. doi: 10.1080/10522150802654278 Rane, T. R., & McBride, B. A. (2000). Identity theory as a guide to understanding fathers' involvement with their children. Journal of Family Issues, 21(3), 347–366. doi: 10.1177/019251300021003004 Rankin, B. H., & Quane, J. M. (2002). Social contexts and urban adolescent outcomes: The interrelated effects of neighborhoods, families, and peers on African-American youth. Social Problems, 49(1), 79–100. doi: 10.1525/sp.2002.49.1.79 Ranote, S., Elliott, R., Abel, K. M., Mitchell, R., Deakin, J. F., & Appleby, L. (2004). The neural basis of maternal responsiveness to infants: An fMRI study. NeuroReport, 15(11), 1825–1829. doi: 10.1097/01.wnr.0000137078.64128.6a Rapee, R. M., Kennedy, S. J., Ingram, M., Edwards, S. L., & Sweeney, L. (2010). Altering the trajectory of anxiety in at-risk young children. The American Journal of Psychiatry, 167, 1518–1525. doi: 10.1176/appi.ajp.2010.09111619 Ratner, N. B., & Pye, C. (1984). Higher pitch in BT is not universal: Acoustic evidence from Quiche Mayan. Journal of Child Language, 11(3), 512–522. Reburn, C. J., & Wynne-Edwards, K. E. (1999). Hormonal changes in males of a naturally biparental and a uniparental mammal. Hormones and Behavior, 35, 163–176. Reichenberg, A., Gross, R., Weisner, M., Bresnahan, M., Silverman, J., Harlap, S.,… Susser, E. (2006). Advancing paternal age and autism. Archives of General Psychiatry, 63(9), 1026– 1032. doi: 10.1001/archpsyc.63.9.1026 Reiss, D. (1995). Genetic influence on family systems: Implications for development. Journal of Marriage and the Family, 57(3), 543–560. Reitman, D., Rhode, P. C., Hupp, S. D. A., & Altobello, C. (2002). Development and

validation of the Parental Authority Questionnaire—Revised. Journal of Psychopathology and Behavioral Assessment, 24(2), 119–127. doi: 10.1023/A:1015344909518 Reitz, E., Deković, M., & Meijer, A. M. (2006). Relations between parenting and externalizing and internalizing problem behaviour in early adolescence: Child behaviour as moderator and predictor. Journal of Adolescence, 29(3), 419–436. doi: 10.1016/j.adolescence.2005.08.003 Repetti, R. L. (1994). Short-term and long-term processes linking job stressors to father–child interaction. Social Development, 3(1), 1–15. Repetti, R. L., Robles, T. F., Reynolds, B. M., & Sears, M. S. (2012). A naturalistic approach to the study of parenting. Parenting: Science and Practice, 12(2–3), 165–174. doi: 10.1080/15295192.2012.683343 Repetti, R. L., & Wang, S. W. (2014). Employment and parenting. Parenting: Science and Practice, 14(2), 121–132. doi: 10.1080/15295192.2014.914364 Repetti, R. L., & Wood, J. (1997). Effects of daily stress at work on mothers' interactions with preschoolers. Journal of Family Psychology, 11(1), 90. Rhoades, K. A., & O'Leary, S. G. (2007). Factor structure and validity of the Parenting Scale. Journal of Clinical Child and Adolescent Psychology, 36(2), 137–146. doi: 10.1080/15374410701274157 Ribas Jr., R. de C., & Bornstein, M. H. (2005). Parenting knowledge: Similarities and differences in Brazilian mothers and fathers. Interamerican Journal of Psychology, 39, 5–12. Rice, F., Harold, G. T., Boivin, J., van den Bree. M., Hay, D. F., & Thapar, A., (2010). The links between prenatal stress and offspring development and psychopathology: Disentangling environmental and inherited influences. Psychological Medicine, 40, 335–345. Richman, A. L., LeVine, R. A., Staples New, R., Howrigan, G. A., Welles-Nystron, B., & LeVine, S. E. (1988). Maternal behavior to infants in five cultures. In R. A. LeVine, P. M. Miller, & M. Maxwell West (Eds.), Parenting behavior in diverse societies (New Directions for Child Development Vol. 40, pp. 81–98). San Francisco, CA: Jossey-Bass. Richman, A. L., Miller, P. M., & LeVine, R. A. (1992). Cultural and educational variations in maternal responsiveness. Developmental Psychology, 28, 614–621. Rilling, J. K. (2013). The neural and hormonal bases of human parental care. Neuropsychologia, 51(4), 731–747. doi: 10.1016/j.neuropsychologia.2012.12.017 Rindfuss, R. R., & Bumpass, L. L. (1978). Age and the sociology of fertility: How old is too old? In K. E. Taeuber, L. L. Bumpass, & J. A. Sweet (Eds.), Social demography (pp. 43–56). New York, NY: Academic Press. Ritter, J. M., Casey, R. J., & Langlois, J. H. (1991). Adults' responses to infants varying in appearance of age and attractiveness. Child Development, 62(1), 68–82.

Roche, K. M., Ensminger, M. E., & Cherlin, A. J. (2007). Variations in parenting and adolescent outcomes among African American and Latino families living in low-income, urban areas. Journal of Family Issues, 28(7), 882–909. doi: 10.1177/0192513X07299617 Roche, K. M., Ghazarian, S. R., Little, T. D., & Leventhal, T. (2011). Understanding links between punitive parenting and adolescent adjustment: The relevance of context and reciprocal associations. Journal of Research on Adolescence, 21(2), 448–460. doi: 10.1111/j.1532– 7795.2010.00681.x Roche, K. M., & Leventhal, T. (2009). Beyond neighborhood poverty: Family management, neighborhood disorder, and adolescents' early sexual onset. Journal of Family Psychology, 23(6), 819. doi: 10.1037/a0016554 Roche, K. M., Mekos, D., Alexander, C. S., Astone, N. M., Bandeen-Roche, K., & Ensminger, M. E. (2005). Parenting influences on early sex initiation among adolescents: How neighborhood matters. Journal of Family Issues, 26(1), 32–54. doi: 10.1177/0192513X04265943 Rodriguez, C. M., & Green, A. J. (1997). Parenting stress and anger expression as predictors of child abuse potential. Child Abuse & Neglect, 21, 367–377. Retrieved from http://search.proquest.com/docview/78965193?accountid=14182 Roe, R. A. (2002). What makes a competent psychologist? European Psychologist, 3, 192– 202. doi: 10.1027//1016–9040.7.3.192 Roksa, J., & Potter, D. (2011). Parenting and academic achievement. Sociology of Education, 84(4), 299–321. Romer, D., Jamieson, P. E., Bushman, B. J., Bleakley, A., Wang, A., Langleben, D., & Jamieson, K. H. (2014). Parental desensitization to violence and sex in movies. Pediatrics, 134(5), 877–884. doi: 10.1542/peds.2014–1167 Rose, R. (1995). Genes and human behavior. Annual Review of Psychology, 46(1), 625–654. Rosenthal, M. K., & Roer-Strier, D. (2001). Cultural differences in mothers' developmental goals and ethnotheories. International Journal of Psychology, 36(1), 20–31. doi: 10.1080/00207590042000029 Rossi, A. S. (1980). Life-span theories and women's lives. Signs: Journal of Women in Culture and Society, 6(1), 4–32. Rothbart, M. K. (2011). Becoming who we are. New York, NY: Guilford Press. Rowe, M. L., Pan, B. A., & Ayoub, C. (2005). Predictors of variation in maternal talk to children: A longitudinal study of low-income families. Parenting: Science and Practice, 5(3), 259–283. doi: 10.1207/s15327922par0503_3

Rubin, K. H., Nelson, L. J., Hastings, P., & Asendorpf, J. (1999). The transaction between parents' perceptions of their children's shyness and their parenting styles. International Journal of Behavioral Development, 23(4), 937–957. Rubin, J. Z., Provenzano, F. J., & Luria, Z. (1974). The eye of the beholder: Parents' views on sex of newborns. American Journal of Orthopsychiatry, 44(4), 512. Ruble, D. N., Martin, C. L., & Berenbaum, S. A. (2006). Gender development. In D. Kuhn, & R. S. Siegler (Ed.), W. Damon (Series Ed.), Handbook of child psychology: Vol. 2. Cognition, Perception, and language (6e, pp. 858–932). Hoboken, NJ: Wiley. Rueter, M. A., & Conger, R. D. (1998). Reciprocal influences between parenting and adolescent problem-solving behavior. Developmental Psychology, 34(6), 1470–1482. Ruiz, M. A., Pincus, A. L., & Schinka, J. A. (2008). Externalizing pathology and the five-factor model: A meta-analysis of personality traits associated with antisocial personality disorder, substance use disorder, and their co-occurrence. Journal of Personality Disorders, 22(4), 365–388. doi: 10.1521/pedi.2008.22.4.365 Ruoppila, I. (1991). The significance of grandparents for the formation of family relations. In P. K. Smith (Ed.), The psychology of grandparenthood: An interactional perspective (pp. 123–139). London, UK: Routledge. Ryan, R. M., Martin, A., & Brooks-Gunn, J. (2006). Is one good parent good enough? Patterns of mother and father parenting and child cognitive outcomes at 24 and 36 months. Parenting: Science and Practice, 6(2–3), 211–228. doi: 10.1080/15295192.2006.9681306 Sadler, L. S., Slade, A., Close, N., Webb, D. L., Simpson, T., Fennie, K., & Mayes, L. C. (2013). Minding the baby: Enhancing reflectiveness to improve early health and relationship outcomes in an interdisciplinary home-visiting program. Infant Mental Health Journal, 34(5), 391–405. doi: 10.1002/imhj.21406 Saigal, S., & Doyle, L. W. (2008). An overview of mortality and sequelae of preterm birth from infancy to adulthood. Lancet, 371(9608), 261–269. doi: 10.1016/S0140– 6736(08)60136–1 Sameroff, A. E. (2009). The transactional model of development: How children and contexts shape each other. Washington, DC: American Psychological Association. doi: 10.1037/11877–000 Sandberg, S. (2013). Lean in: Women, work, and the will to lead. New York, NY: Knopf. Sanders, M. (2010). Adopting a public health approach to the delivery of evidence-based parenting interventions. Canadian Psychology, 51(1), 17–23. doi: 10.1037/a0018295 Sanders, M. R., & Kirby, J. N. (2010). Parental programs for preventing behavioral and emotional problems in children. In J. Bennet-Levy, D. Richards, P. Farrand, H. Christensen, K.

Griffiths, D. Kavanagh,… C. Williams (Eds.), Oxford guide to low intensity CBT interventions (pp. 399–406). New York, NY: Oxford University Press. Sanders, M. R., Markie-Dadds, C., & Turner, K. M. (2003). Theoretical, scientific and clinical foundations of the Triple P-Positive Parenting Program: A population approach to the promotion of parenting competence (Vol. 1). Queensland, Australia: Parenting and Family Support Centre, University of Queensland. Sandler, I., Schoenfelder, E., Wolchik, S., & MacKinnon, D. (2011). Long-term impact of prevention programs to promote effective parenting: Lasting effects but uncertain processes. Annual Review of Psychology, 62, 299–329. doi: 10.1146/annurev.psych.121208.131619 Sanghavi, D. M. (2005). Taking well-child care into the 21st century—A novel, effective method for improving parent knowledge using computerized tutorials. Archives of Pediatrics and Adolescent Medicine, 159(5), 482–485. doi: 10.1001/archpedi.159.5.482 Sargent, J. (1983). The sick child: Family complications. Developmental and Behavioral Pediatrics, 4(1), 50–56. Sawhill, I. V. (2014). Generation unbound: Drifting into sex and parenthood without marriage. Washington, DC: Brookings Institution Press. Saxbe, D. E., Repetti, R. L., & Nishina, A. (2008). Marital satisfaction, recovery from work, and diurnal cortisol among men and women. Health Psychology, 27(1), 15. doi: 10.1037/0278–6133.27.1.15 Sayer, L., Bianchi, S., & Robinson, J. (2004). Are parents investing less in children? Trends in mothers' and fathers' time with children. American Journal of Sociology, 110(1), 1–43. doi: 10.1086/386270 Scaramella, L. V., Neppl, T. K., Ontai, L. L., & Conger, R. D. (2008). Consequences of socioeconomic disadvantage across three generations: Parenting behavior and child externalizing problems. Journal of Family Psychology, 22(5), 725. doi: 10.1037/a0013190 Scarr, S., & Kidd, K. K. (1983). Developmental behavior genetics. In M. M. Haith & J. J. Campos (Vol. Eds.), Infancy and developmental psychobiology. In P. H. Mussen (Editor-inChief), Handbook of child psychology (4th ed., Vol. 2, pp. 345–433). New York, NY: Wiley. Schaller, J. (2012). Booms, busts, and fertility: Testing the Becker model using genderspecific labor demand. Unpublished manuscript, University of Arizona. Scheers, N. J., Rutherford, G. W., & Kemp, J. S. (2003). Where should infants sleep? A comparison of risk for suffocation of infants sleeping in cribs, adult beds, and other sleeping locations. Pediatrics, 112(4), 883–889. doi: 10.1542/peds.112.4.883 Scheper-Hughes, N. (1989). Human strategy: Death without weeping. Natural History Magazine, 98(10), 8–16.

Schlomer, G. L., & Belsky, J. (2012). Maternal age, investment, and parent-child conflict: A mediational test of the terminal investment hypothesis. Journal of Family Psychology, 26, 443–452. doi: 10.1037/a0027859 Schoppe-Sullivan, S. J., Brown, G. L., Cannon, E. A., Mangelsdorf, S. C., & Sokolowski, M. S. (2008). Maternal gatekeeping, coparenting quality, and fathering behavior in families with infants. Journal of Family Psychology, 22(3), 389–398. doi: 10.1037/0893–3200.22.3.389 Schreier, H. M., & Chen, E. (2012). Socioeconomic status and the health of youth: A multilevel, multidomain approach to conceptualizing pathways. Psychological Bulletin, 139(3), 606–654. Schulze, P., Harwood, R., & Schoelmerich, A. (2001). Developmental expectations and feeding practices of Puerto Rican and Anglo mothers. Journal of Cross-Cultural Psychology, 32, 397–406. Scrimin, S., Haynes, O. M., Altoè, G., Bornstein, M. H., & Axia, G. (2009). Anxiety and stress in mothers and fathers in the 24 h after their child's surgery. Child: Care, Health and Development, 35, 227–233. Seavey, C. A., Katz, P. A., & Zalk, S. R. (1975). Baby X. Sex Roles, 1(2), 103–109. Seip, K. M., & Morrell, J. I. (2007). Increasing the incentive salience of cocaine challenges preference for pup-over cocaine-associated stimuli during early postpartum: Place preference and locomotor analyses in the lactating female rat. Psychopharmacology, 194(3), 309–319. doi: 10.1007/s00213–007–0841–9 Selin, H., & Stone, P. K. (Eds.). (2009). Childbirth across cultures: Ideas and practices of pregnancy, childbirth, and the postpartum. Dordrecht, the Netherlands: Springer Science + Business Media. Senese, V. P., Bornstein, M. H., Haynes, O. M., Rossi, G., & Venuti, P. (2012). A cross-cultural comparison of mothers' beliefs about their parenting very young children. Infant Behavior and Development, 35(3), 479–488. doi: 10.1016/j.infbeh.2012.02.006 Senese, V. P., de Falco, S., Bornstein, M. H., Caria, A. Buffolino, S., & Venuti, P. (2013). Human infant faces provoke implicit positive affective responses in parents and non parents alike. PLoS ONE, 8, e80379, 1–7. Sevigny, P. R., & Loutzenhiser, L. (2010). Predictors of parenting self-efficacy in mothers and fathers of toddlers. Child: Care, Health and Development, 36(2), 179–189. doi: 10.1111/j.1365–2214.2009.00980.x Shadish, W. R., Cook, T. D., & Campbell, D. T. (2002). Experimental and quasi-experimental designs for generalized causal inference. Boston, MA: Houghton Mifflin. Shah, P. S., Zao, J., & Ali, S. (2011). Maternal marital status and birth outcomes: A systematic

review and meta-analyses. Maternal and Child Health Journal, 15(7), 1097–1109. doi: 10.1007/s10995–010–0654-z Shakin, M., Shakin, D., & Sternglanz, S. H. (1985). Infant clothing: Sex labeling for strangers. Sex Roles, 12(9–10), 955–964. Shapiro, C. J., Prinz, R. J., & Sanders, M. R. (2012). Facilitators and barriers to implementation of an evidence-based parenting intervention to prevent child maltreatment: The Triple P-Positive Parenting Program. Child Maltreatment, 17(1), 86–95. doi: 10.1177/1077559511424774 Shaver, P. R., & Mikulincer, M. (2002). Attachment-related psychodynamics. Attachment & Human Development, 4, 133–161. Shaw, P., Greenstein, D., Lerch, J., Clasen, L., Lenroot, R., Gogtay, N.,… Giedd, J. (2006). Intellectual ability and cortical development in children and adolescents. Nature, 440(7084), 676–679. doi: 10.1038/nature04513 Sheehan, T., & Numan, M. (2002). Estrogen, progesterone, and pregnancy termination alter neural activity in brain regions that control maternal behavior in rats. Neuroendocrinology, 75(1), 12–23. doi: 10.1159/000048217 Shumow, L., & Lomax, R. (2002). Parental efficacy: Predictor of parenting behavior and adolescent outcomes. Parenting: Science and Practice, 2(2), 127–150. doi: 10.1207/S15327922PAR0202_03 Siddiqui, A., & Hägglöf, B. (2000). Does maternal prenatal attachment predict postnatal mother infant interaction? Early Human Development, 59(1), 13–25. doi: 10.1016/S0378– 3782(00)00076–1 Sidorowicz, L. S., & Lunney, G. S. (1980). Baby X revisited. Sex Roles, 6(1), 67–73. Sigel, I. E., & McGillicuddy-De Lisi, A. V. (2002). Parental beliefs and cognitions: The dynamic belief systems model. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 485–508). Mahwah, NJ: Erlbaum. Silver, M. (1965). Births, marriages, and business cycles in the United States. Journal of Political Economics, 73(3), 237–255. Simerly, R. B. (2002). Wired for reproduction: Organization and development of sexually dimorphic circuits in the mammalian forebrain. Annual Review of Neuroscience, 25(1), 507– 536. doi: 10.1146/annurev.neuro.25.112701.142745 Simons, R. L., Murry, V., McLoyd, V., Lin, K. H., Cutrona, C., & Conger, R. D. (2002). Discrimination, crime, ethnic identity, and parenting as correlates of depressive symptoms among African American children: A multilevel analysis. Development and Psychopathology, 14(02), 371–393. doi: 10.1017/S0954579402002109

Simons, R. L., Simons, L. G., Burt, C. H., Brody, G. H., & Cutrona, C. (2005). Collective efficacy, authoritative parenting and delinquency: A longitudinal test of a model integrating community- and family-level processes. Criminology, 43(4), 989–1029. doi: 10.1111/j.1745– 9125.2005.00031.x Skinner, A. T., Bacchini, D., Lansford, J. E., Godwin, J., Sorbring, E., Tapanya, S.,… Pastorelli, C. (2014). Neighborhood danger, parental monitoring, harsh parenting, and child aggression in nine countries. Societies, 4, 45–67. Skinner, A. T., Oburu, P., Lansford, J. E., & Bacchini, D. (2014). Childrearing violence and child adjustment after exposure to Kenyan post-election violence. Psychology of Violence, 4(1), 37–50. doi: 10.1037/a0033237 Skinner, E., Johnson, S., & Snyder, T. (2005). Six dimensions of parenting: A motivational model. Parenting: Science and Practice, 5(2), 175–235. doi: 10.1207/s15327922par0502_3 Slep, A. M. S., & O'Leary, S. G. (1998). The effects of maternal attributions on parenting: An experimental analysis. Journal of Family Psychology, 12(2), 234. Smetana, J. G. (2000). Middle-class African American adolescents' and parents' conceptions of parental authority and parenting practices: A longitudinal investigation. Child Development, 71(6), 1672–1686. doi: 10.1111/1467–8624.00257 Smetana, J. G., & Villalobos, M. (2009). Social cognitive development in adolescence. In R. M. Lerner & L. Steinberg (Eds.), Handbook of adolescent psychology (pp. 187–229). Hoboken, NJ: Wiley. doi: 10.1002/9780470479193.adlpsy001008 Smith, C. A., & Farrington, D. P. (2004). Continuities in antisocial behavior and parenting across three generations. Journal of Child Psychology and Psychiatry, 45(2), 230–247. doi: 10.1111/j.1469–7610.2004.00216.x Smith, C. L., Spinrad, T. L., Eisenberg, N., Gaertner, B. M., Popp, T. K., & Maxon, E. (2007). Maternal personality: Longitudinal associations to parenting behavior and maternal emotional expressions toward toddlers. Parenting: Science and Practice, 7(3), 305–329. doi: 10.1080/15295190701498710 Smith, P. K., & Drew, L. M. (2002). Grandparenthood. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 141–172). Mahwah, NJ: Erlbaum. Snell, W. E., Overbey, G., & Brewer, A. L. (2005). Parenting perfectionism and the parenting role. Personality and Individual Differences, 39(3), 613–624. doi: 10.1016/j.paid.2005.02.006 Snider, J. B., Clements, A., & Vazsonyi, A. T. (2004). Late adolescent perceptions of parent religiosity and parenting processes. Family Process, 43(4), 489–502. doi: 10.1111/j.1545– 5300.2004.00036.x

Snow, C. E. (1977). Mothers' speech research: From input to interactions. In C. E. Snow & C. A. Ferguson (Eds.), Talking to children: Language input and acquisition (pp. 31–49). London, UK: Cambridge University Press. Sobotka, T., Skirbekk, V., & Philipov, D. (2011) Economic recession and fertility in the developed world. Population Development Review, 37(2), 267–306. Soderstrom, M. (2007). Beyond babytalk: Re-evaluating the nature and content of speech input to preverbal infants. Developmental Review, 27(4), 501–532. doi: 10.1016/j.dr.2007.06.002 Soenens, B., Luyckx, K., Vansteenkiste, M., Duriez, B., & Goossens, L. (2008). Clarifying the link between parental psychological control and adolescents' depressive symptoms. MerrillPalmer Quarterly, 54(4), 411–444. Solmeyer, A. R., Killoren, S. E., McHale, S. M., & Updegraff, K. A. (2011). Coparenting around siblings' differential treatment in Mexican-origin families. Journal of Family Psychology, 25, 251–260. doi: 10.1037/a0023201 Solomon, A. (2012). Far from the tree: Parents, children, and the search for identity. New York, NY: Scribner. Soltis, J. (2004). The signal functions of early infant crying. Behavioral and Brain Sciences, 27(4), 443–458. doi: 10.1017/S0140525X0400010X Sorensen, E. (1997). A national profile of nonresident fathers and their ability to pay child support. Journal of Marriage and the Family, 59(4), 785–797. Spitz, R. A. (1970). The first year of life. New York, NY: International Universities Press. (Original work published 1965) Spock, B., & Needlman, R. (2013). Dr. Spock's baby and child care (9th ed.). New York, NY: Simon & Schuster. Sprengelmeyer, R., Perrett, D. I., Fagan, E. C., Cornwell, R. E., Lobmaier, J. S., Sprengelmeyer, A.,… Young, A. W. (2009). The cutest little baby face: A hormonal link to sensitivity to cuteness in infant faces. Psychological Science, 20(2), 149–154. doi: 10.1111/j.1467–9280.2009.02272.x Sroufe, A. (1985). Attachment classification from the perspective of infant-caregiver relationships and infant temperament. Child Development, 56, 1–14. Sroufe, L. A., & Jacobvitz, D. (1989). Diverging pathways, developmental transformations, multiple etiologies and the problem of continuity in development. Human Development, 32, 196–203. Stallings, J., Fleming, A. S., Corter, C., Worthman, C., & Steiner, M. (2001). The effects of infant cries and odors on sympathy, cortisol, and autonomic responses in new mothers and nonpostpartum women. Parenting: Science and Practice, 1(1–2), 71–100. doi:

10.1080/15295192.2001.9681212 Stattin, H., & Kerr, M. (2000). Parental monitoring: A reinterpretation. Child Development, 71(4), 1072–1085. doi: 10.1111/1467–8624.00210 Stein, A., Craske, M. G., Lehtonen, A., Harvey, A., Savage-McGlynn, E., Davies, B.,… Counsell, N. (2012). Maternal cognitions and mother-infant interaction in postnatal depression and generalized anxiety disorder. Journal of Abnormal Psychology, 121(4), 795–809. doi: 10.1037/a0026847 Steinberg, L., & Silk, J. S. (2002). Parenting adolescents. In M. H. Bornstein (Ed.), Handbook of parenting Vol. 1 Children and parenting (2e, pp. 103–133). Mahwah, NJ: Erlbaum. Stelle, C., Fruhauf, C. A., Orel, N., & Landry-Meyer, L. (2010). Grandparenting in the 21st century: Issues of diversity in grandparent–grandchild relationships. Journal of Gerontological Social Work, 53(8), 682–701. doi: 10.1080/01634372.2010.516804 Stephan, C. W., & Langlois, J. H. (1984). Baby beautiful: Adult attributions of infant competence as a function of infant attractiveness. Child Development, 55, 576–585. Stice, E., & Barrera, M. (1995). A longitudinal examination of the reciprocal relations between perceived parenting and adolescents' substance use and externalizing behaviors. Developmental Psychology, 31(2), 322. Stith, S. M., Rosen, K. H., Middleton, K. A., Busch, A. L., Lundeberg, K., & Carlton, R. P. (2000). The intergenerational transmission of spouse abuse: A meta-analysis. Journal of Marriage & the Family, 62(3), 640–654. doi: 10.1111/j.1741–3737.2000.00640.x St. James-Roberts, S. (2008). Infant crying and sleeping: helping parents to prevent and manage problems. Primary Care: Clinics in Office Practice, 35(3), 547–567. doi: 10.1016/j.pop.2008.06.004 Stolz, H. E., Barber, B. K., & Olsen, J. A. (2005). Toward disentangling fathering and mothering: An assessment of relative importance. Journal of Marriage and Family, 67(4), 1076–1092. doi: 10.1111/j.1741–3737.2005.00195.x Storey, A. E., Walsh, C. J., Quinton, R. L., & Wynne-Edwards, K. E. (2000). Hormonal correlates of paternal responsiveness in new and expectant fathers. Evolution and Human Behavior, 21(2), 79–95. doi: 10.1016/S1090–5138(99)00042–2 Stormshak, E. A., Bierman, K. L., McMahon, R. J., & Lengua, L. J. (2000). Parenting practices and child disruptive behavior problems in early elementary school. Journal of Clinical Child Psychology, 29(1), 17–29. doi: 10.1207/S15374424jccp2901_3 Strassberg, Z., & Treboux, D. (2000). Interpretations of child emotion expressions and coercive parenting practices of adolescent mothers. Social Development, 9(1), 80–95. doi: 10.1111/1467–9507.00112

Strassman, B. I., & Gillespie, B. (2002). Life-history theory, fertility and reproductive success in humans. Proceedings of the Royal Society of London, 269, 553–562. doi: 10.1098/rspb.2001.1912 Strathearn, L., Li, J., Fonagy, P., & Montague, P. R. (2008). What's in a smile? Maternal brain responses to infant facial cues. Pediatrics, 122(1), 40–51. doi: 10.1542/peds.2007–1566 Strauss, S., & Stavey, R. (1982). U-shaped behavioral growth: Implications for theories of development. In W. W. Hartup (Ed.), Review of child development research (Vol. 6, pp. 547– 599). Chicago, IL: University of Chicago Press. Stright, A. D., & Bales, S. S. (2003). Coparenting quality: Contributions of child and parent characteristics. Family Relations, 52(3), 232–240. doi: 10.1111/j.1741–3729.2003.00232.x Stuckey, F., McGhee, P. E., & Bell, N. J. (1982). Parent-child interaction: The influence of maternal employment. Developmental Psychology, 18, 635–644. Suchman, N. E., & Luthar, S. S. (2001). The mediating role of parenting stress in methadone maintained mothers' parenting. Parenting: Science and Practice, 1(4), 285–315. doi: 10.1207/S15327922PAR0104_2 Suitor, J. J., Sechrist, J., Plikuhn, M., Pardo, S. T., Gilligan, M., & Pillemer, K. (2009). The role of perceived maternal favoritism in sibling relations in midlife. Journal of Marriage and Family, 71(4), 1026–1038. doi: 10.1111/j.1741–3737.2009.00650.x Suizzo, M. A., & Cheng, C. C. (2007). Taiwanese and American mothers' goals and values for their children's futures. International Journal of Psychology, 42(5), 307–316. doi: 10.1080/00207590601109342 Sulloway, F. J. (1996). Born to rebel: Birth order, family dynamics, and creative lives. New York, NY: Vintage. Susser, E., Hoek, H. W., & Brown, A. (1998). Neurodevelopmental disorders after prenatal famine. American Journal of Epidemiology, 147(3), 213–216. Suzuki, S., Holloway, S. D., Yamamoto, Y., & Mindnich, J. D. (2009). Parenting self-efficacy and social support in Japan and the United States. Journal of Family Issues, 30(11), 1505– 1526. doi: 10.1177/0192513X09336830 Swain, J. E., Leckman, J. F., Mayes, L. C., Feldman, R., & Schultz, R. T. (2006). Own baby pictures induce parental brain activations according to psychology, experience and postpartum timing. Biological Psychiatry, 59, 126S. Swain, J. E., Lorberbaum, J. P., Kose, S., & Strathearn, L. (2007). Brain basis of early parent infant interactions: Psychology, physiology, and in vivo functional neuroimaging studies. Journal of Child Psychology and Psychiatry, 48(3–4), 262–287. doi: 10.1111/j.1469– 7610.2007.01731.x

Szyf, M., & Bick, J. (2013). DNA methylation: A mechanism for embedding early life experiences in the genome. Child Development, 84(1), 49–57. doi: 10.1111/j.1467– 8624.2012.01793.x Takanishi, R., & Bogard, K. L. (2007). Effective and educational programs for young children: What we need to know. Child Development Perspectives, 1(1), 40–45. doi: 10.1111/j.1750– 8606.2007.00008.x Tamis-LeMonda, C. S., & Cabrera, N. (Eds.). (2002). Handbook of father involvement: Multidisciplinary perspectives. Mahwah, NJ: Erlbaum. Tamis-LeMonda, C. S., & Kahana-Kalman, R. (2009). Mothers' views at the transition to a new baby: Variation across ethnic groups. Parenting: Science and Practice, 9, 35–55. doi: 10.1080/15295190802656745 Tamis-LeMonda, C. S., & McFadden, K. E. (2010). The United States of America. In M. H. Bornstein (Ed.), Handbook of cultural developmental science (pp. 299–322). New York, NY: Psychology Press. Tamis-LeMonda, C. S., Shannon, J. D., Cabrera, N. J., & Lamb, M. E. (2004). Fathers and mothers at play with their 2- and 3-year-olds: Contributions to language and cognitive development. Child Development, 75(6), 1806–1820. doi: 10.1111/j.1467– 8624.2004.00818.x Tamis-LeMonda, C. S., Wang, S., Koutsouvanou, E., & Albright, M. (2002). Childrearing values in Greece, Taiwan, and the United States. Parenting: Science and Practice, 2(3), 185– 208. doi: 10.1207/S15327922PAR0203_01 Tan, T., Camras, L. A., Deng, H., Zhang, M., & Lu, Z. (2012). Family stress, parenting styles, and behavioral adjustment in preschool-age adopted Chinese girls. Early Childhood Research Quarterly, 27(1), 128–136. doi: 10.1016/j.ecresq.2011.04.002 Tanskanen, A. O. & Rotkirch, A. (2014). The impact of grandparental investment on mothers' fertility intentions in four European countries. Demographic Research, 30, 1–26. Taylor, C. A., Guterman, N. B., Lee, S. J., & Rathouz, P. J. (2009). Intimate partner violence, maternal stress, nativity, and risk for maternal maltreatment of young children. American Journal of Public Health, 99(1), 175–183. doi: 10.2105/AJPH.2007.126722 Teti, D. M., & Candelaria, M. (2002). Parenting competence. In M. H. Bornstein (Ed.), Handbook of parenting: Applied parenting (2nd ed., Vol. 4, pp. 149–180). Mahwah, NJ: Erlbaum. Teti, D. M., & Gelfand, D. M. (1991). Behavioral competence among mothers of infants in the first year: the mediational role of maternal self-efficacy. Child Development, 62(5), 918–929. Teubert, D., & Pinquart, M. (2010). The association between coparenting and child

adjustment: A meta-analysis. Parenting: Science and Practice, 10(4), 286–307. doi: 10.1080/15295192.2010.492040 Thompson, E., & Miller, P. C. (1997, April). Parental beliefs and use of parental discipline: The role of religious affiliation. Poster presented at the biennial meeting of the Society for Research in Child Development, Washington, DC. Thornberry, T. P. (2005). Explaining multiple patterns of offending across the life course and across generations. ANNALS, American Academy of Political and Social Science, 602(1), 156–194. doi: 10.1177/0002716205280641 Thornberry, T. P., Freeman-Gallant, A., Lizotte, A. J., & Krohn, M. D. (2003). Linked lives: The intergenerational transmission of antisocial behavior. Journal of Abnormal Child Psychology, 31(2), 171–184. Thun-Hohenstein, L., Wienerroither, C., Schreuer, M., Seim, G., & Wienerroither, H. (2008). Antenatal mental representations about the child and mother–infant interaction at three months post partum. European Child & Adolescent Psychiatry, 17(1), 9–19. doi: 10.1007/s00787– 007–0622–3 Tinbergen, N. (1963). On aims and methods of ethology. Zeitschrift für Tierpsychologie, 20(4), 410–433. Toldson, I. A., & Lemmons, B. P. (2013). Social demographics, the school environment, and parenting practices associated with parents' participation in schools and academic success among Black, Hispanic, and White students. Journal of Human Behavior in the Social Environment, 23(2), 237–255. doi: 10.1080/10911359.2013.747407 Tomlinson, M., Bornstein, M. H., Marlow, M., & Swartz, L. (2014). Imbalances in the knowledge about infant mental health in rich and poor countries: Too little progress in bridging the gap. Infant Mental Health Journal, 35(6), 624–629. doi: 10.1002/imhj.21462 Trainor, B. C., & Marler, C. A. (2001). Testosterone, paternal behavior, and aggression in the monogamous California mouse (Peromyscus californicus). Hormones and Behavior, 40(1), 32–42. doi: 10.1006/hbeh.2001.1652 Trainor, B. C., & Marler, C. A. (2002). Testosterone promotes paternal behaviour in a monogamous mammal via conversion to oestrogen. Proceedings of the Royal Society of London, Series B: Biological Sciences, 269(1493), 823–829. doi: 10.1098/rspb.2001.1954 Trevarthen, C. (1979). Communication and cooperation in early infancy: A description of primary intersubjectivity. In M. Bullowa (Ed.), Before speech: The beginning of interpersonal communication (pp. 321–347). New York, NY: Cambridge University Press. Treyvaud, K., Rogers, S., Matthews, J., & Allen, B. (2010). Maternal factors and experiences associated with observed parenting behavior in mothers attending a residential parenting program. Infant Mental Health Journal, 31(1), 58–70. doi: 10.1002/imhj.20242

Triandis, H. C. (1995). Individualism and collectivism. Boulder, CO: Westview Press. Trinder, L. (2008). Maternal gate closing and gate opening in postdivorce families. Journal of Family Issues, 29(10), 1298–1324. doi: 10.1177/0192513X08315362 Trivers, R. L. (1972). Parental investment and sexual selection. In B. Campbell (Ed.), Sexual selection and the descent of man, 1871–1971 (pp. 136–179). Chicago, IL: Aldine. Trumbo, S. T. (1996). Parental care in invertebrates. Advances in the Study of Behavior, 25, 351. Tudge, J., Hogan, D., Lee, S., Tammeveski, P., Meltsas, M., Kulakova, N., Snezhkova, I. & Putnam, S. (1999). Cultural heterogeneity: Parental values and beliefs and their preschoolers' activities in the United States, South Korea, Russia, and Estonia. In A. Göncü (Ed.), Children's engagement in the world. Sociocultural perspectives (pp. 62–96). Cambridge: Cambridge University Press. Tuffrey, C., & Finlay, F. (2002). Use of the Internet by parents of paediatric outpatients. Archives of Disease in Childhood, 87(6), 534–536. doi: 10.1136/adc.87.6.534 Tulviste, T., Mizera, L., De Geer, B., & Tryggvason, M. T. (2007). Child-rearing goals of Estonian, Finnish, and Swedish mothers. Scandinavian Journal of Psychology, 48(6), 487– 497. Turkheimer, E. (1998). Heritability and biological explanation. Psychological Review, 105(4), 1–10. Turkheimer, E., & Waldron, M. (2000). Nonshared environment: A theoretical, methodological, and quantitative review. Psychological Bulletin, 126(1), 78. doi: 10.1037/0033–2909.126.1.78 Ungar, M., Ghazinour, M., & Richter, J. (2013). Annual research review: What is resilience within the social ecology of human development? Journal of Child Psychology and Psychiatry, 54(4), 348–366. doi: 10.1111/jcpp.12025 UNICEF. (2006). The state of the world's children 2007: Women and children: The double dividend of gender equality (Vol. 7). New York City: UNICEF. United States Bureau of Labor Statistics. (2013). American time use survey—2012. Retrieved from http://www.bls.gov/news.release/pdf/atus.pdf United States Department of Labor. (2011). Women's employment during the recovery. Retrieved from http://www.dol.gov/_sec/media/reports/FemaleLaborForce/FemaleLaborForce.pdf Updegraff, K. A., McHale, S. M., Crouter, A. C., & Kupanoff, K. (2001). Parents' involvement in adolescents' peer relationships: A comparison of mothers' and fathers' roles. Journal of Marriage and Family, 63(3), 655–668. doi: 10.1111/j.1741–3737.2001.00655.x

Valsiner, J. (2000). Culture and human development: An introduction. London: Sage. van Anders, S., Tolman, R. M., & Volling, B. L. (2012). Baby cries and nurturance affect testosterone in men. Hormones and Behavior, 61(1), 31–36. doi: 10.1016/j.yhbeh.2011.09.012 Van den Akker, A. L., Deković, M., Asscher, J., & Prinzie, P. (2014). Mean-level personality development across childhood and adolescence: A temporary defiance of the maturity principle and bidirectional associations with parenting. Journal of Personality and Social Psychology, 107(4), 736–750. van den Boom, D. C. (1994). The influence of temperament and mothering on attachment and exploration: An experimental manipulation of sensitive responsiveness among lower-class mothers with irritable infants. Child Development, 65(5), 1457–1477. van den Bloom, D. C., & Hoeksma, J. B. (1994). The effect of infant irritability on motherinfant interaction: A growth-curve analysis. Developmental Psychology, 30(4), 581–590. doi: 10.1037/0012–1649.30.4.581 van der Bruggen, C. O., Stams, G. J. J. M., & Bogels, S. M. (2008). The relationship between children and parent anxiety and parental control: A meta-analytic review. Journal of Child Psychology and Psychiatry, 49(12), 1257–1269. doi: 10.1111/j.1469–7610.2008.01898.x van Doesum, K. T. M., Hosman, C. M., Riksen-Walraven, J., & Hoefnagels, C. (2007). Correlates of depressed mothers' sensitivity toward their infants: The role of maternal, child, and contextual characteristics. Journal of the American Academy of Child & Adolescent Psychiatry, 46(6), 747–756. doi: 10.1097/CHI.0b013e318040b272 van Doesum, K. T. M., Riksen-Walraven, J. M., Hosman, C. M. H., & Hoefnagels, C. (2008). A randomized controlled trial of a home-visiting intervention aimed at preventing relationship problems in depressed mothers and their infants. Child Development, 79(3), 547–561. doi: 10.1111/j.1467–8624.2008.01142.x Van Egeren, L. A., & Hawkins, D. P. (2004). Coming to terms with coparenting: Implications of definition and measurement. Journal of Adult Development, 11(3), 165–178. doi: 10.1023/B:JADE.0000035625.74672.0b van IJzendoorn, M. H. (1992). Intergenerational transmission of parenting: A review of studies in nonclinical populations. Developmental Review, 12(1), 76–99. van IJzendoorn, M. H. (1995). Adult attachment representations, parental responsiveness, and infant attachment: A meta-analysis on the predictive validity of the adult attachment interview. Psychological Bulletin, 117(3), 387–403. van IJzendoorn, M. H., Bakermans-Kranenburg, M. J., & Mesman, J. (2008). Dopamine system genes associated with parenting in the context of daily hassles. Genes, Brain, and Behavior,

7(4), 403–410. doi: 10.1111/j.1601–183X.2007.00362.x Vandell, D. L. (2000). Parents, peer groups, and other socializing influences. Developmental Psychology, 36(6), 699–710. doi: 10.1037/0012–1649.36.6.699 Vassoler, F. M., White, S. L., Schmidt, H. D., Sadri-Vakili, G., & Pierce, R. C. (2012). Epigenetic inheritance of a cocaine-resistance phenotype. Nature Neuroscience, 16(1), 42–47. doi: 10.1038/nn.3280 Veenema, A. H. (2009). Early life stress, the development of aggression and neuroendocrine and neurobiological correlates: What can we learn from animal models? Frontiers in Neuroendocrinology, 30(4), 497–518. doi: 10.1016/j.yfrne.2009.03.003 Venuti, P., Caria, A., Esposito, G., de Pisapia, N., Bornstein, M. H., & de Falco, S. (2012). Differential brain responses to cries of infants with autistic disorder and typical development: An fMRI study. Research in Developmental Disabilities, 33(6), 2255–2264. doi: 10.1016/j.ridd.2012.06.011 Verhoeven, M., Junger, M., Van Aken, C., Deković, M., & Van Aken, M. A. G. (2007). Parenting during toddlerhood: Contributions of parental, contextual, and child characteristics. Journal of Family Issues, 28(12), 1663–1691. doi: 10.1177/0192513X07302098 Vermulst, A. A., de Brock, A. J. L. L., & van Zutphen, R. A. H. (1991). Transmission of parenting across generations. In P. K. Smith (Ed.), The psychology of grandparenthood: An interactional perspective (pp. 100–122). London, UK: Routledge. Vernberg, E. M., Beery, S. H., Ewell, K. K., & Absender, D. A. (1993). Parents' use of friendship facilitation strategies and the formation of friendships in early adolescence: A prospective study. Journal of Family Psychology, 7(3), 356. Vesga-López, O., Blanco, C., Keyes, K., Olfson, M., Grant, B. F., & Hasin, D. S. (2008). Psychiatric disorders in pregnant and postpartum women in the United States. Archives of General Psychiatry, 65(7), 805–815. doi: 10.1001/archpsyc.65.7.805 Visscher, P. M., Hill, W. G., & Wray, N. R. (2008). Heritability in the genomics era—concepts and misconceptions. Nature Reviews Genetics, 9(4), 255–266. doi: 10.1038/nrg2322 Volk, A. A., & Quinsey, V. L. (2002). The influence of infant facial cues on adoption preferences. Human Nature, 13(4), 437–455. doi: 10.1007/s12110–002–1002–9 Volling, B. L., Mahoney, A., & Rauer, A. J. (2009). Sanctification of parenting, moral socialization, and young children's conscience development. Psychology of Religion and Spirituality, 1(1), 53. doi: 10.1037/a0014958 von der Lippe, A. L. (1999). The impact of maternal schooling and occupation on child-rearing attitudes and behaviours in low income neighbourhoods in Cairo, Egypt. International Journal of Behavioral Development, 23, 703–729.

Vondra, J., Sysko, H. B., & Belsky, J. (2005). Developmental origins of parenting: Personality and relationship factors. In T. Luster & L. Okagaki (Eds.), Parenting: An ecological perspective (2nd ed., pp. 35–71). Mahwah, NJ: Erlbaum. Vygotsky, L. (1978). Mind in society. Cambridge, MA: Harvard University Press. Wachs, T. D. (2005, April). Environmental chaos and children's development: Expanding the boundaries of chaos. Paper presented at the Society for Research in Child Development, Atlanta, GA. Wachs, T. D., & Camli, O. (1991). Do ecological or individual characteristics mediate the influence of the physical environment upon maternal behavior. Journal of Environmental Psychology, 11, 249–264. Wahlsten, D. (2012). The hunt for gene effects pertinent to behavioral traits and psychiatric disorders: From mouse to human. Developmental Psychobiology, 54(5), 475–492. doi: 10.1002/dev.21043 Waizenhofer, R. N., Buchanan, C. M., & Jackson-Newsom, J. (2004). Mothers' and fathers' knowledge of adolescents' daily activities: Its sources and its links with adolescent adjustment. Journal of Family Psychology, 18(2), 348. doi: 10.1037/0893–3200.18.2.348 Walker, A. J., & McGraw, L. A. (2000). Who is responsible for responsible fathering? Journal of Marriage and Family, 62(2), 563–596. doi: 10.1111/j.1741–3737.2000.00563.x Walker, S. K., Dworkin, J., & Connell, J. (2011). Variation in parent use of information and communications technology: Does quantity matter? Family and Consumer Sciences Research Journal, 40(2), 106–119. doi: 10.1111/j.1552–3934.2011.02098.x Wang, M. T., & Kenny, S. (2014). Longitudinal links between fathers' and mothers' harsh verbal discipline and adolescents' conduct problems and depressive symptoms. Child Development, 85(3), 908–923. doi: 10.1111/cdev.12143 Wang, S., Repetti, R. L., & Campos, B. (2011). Job stress and family social behavior: The moderating role of neuroticism. Journal of Occupational Health Psychology, 16, 441–456. Wang, W., & Parker, K. (2014). Record share of Americans have never married: As values, economics and gender patterns change. Retrieved from Pew Research Center's Social & Demographic Trends website: http://www.pewsocialtrends.org/files/2014/09/2014–09– 24_Never-Married-Americans.pdf Wang, Z., Deater-Deckard, K., & Bell, M. A. (2013). Household chaos moderates the link between maternal attribution bias and parenting. Parenting: Science and Practice, 13(4), 233–252. doi: 10.1080/15295192.2013.832569 Warren, J., Allen, M., Okuyemi, K., Kvasny, L., & Hecht, M. (2010). Targeting single parents in preadolescent substance use prevention: Internet characteristics and information relevance.

Drugs: Education, Prevention, and Policy, 17(4), 400–412. doi: 10.3109/09687630802559083 Way, N., & Greene, M. L. (2006). Trajectories of perceived friendship quality during adolescence: The patterns and contextual predictors. Journal of Research on Adolescence, 16(2), 293–320. doi: 10.1111/j.1532–7795.2006.00133.x Weaver, C. M., Shaw, D. S., Dishion, T. J., & Wilson, M. N. (2008). Parenting self-efficacy and problem behavior in children at high risk for early conduct problems: The mediating role of maternal depression. Infant Behavior & Development, 31(4), 594–605. doi: 10.1016/j.infbeh.2008.07.006 Weaver, I. C. G., Cervoni, N., Champagne, F. A., D'Alessio, A. C., Sharma, S., Seckl, J. R., & Meaney, M. J. (2004). Epigenetic programming by maternal behavior. Nature Neuroscience, 7(8), 847–854. doi: 10.1038/nn1276 Weaver, N. L., Williams, J., Jacobsen, H. A., Botello-Harbaum, M., Glasheen, C., Noelcke, E., & Nansel, T. R. (2008). Translation of an evidence-based tailored childhood injury prevention program. Journal of Public Health Management and Practice, 14(2), 177. doi: 10.1097/01.PHH.0000311897.03573.cc Wee, K. Y., Skouteris, H., Pier, C., Richardson, B., & Milgrom, J. (2011). Correlates of anteand postnatal depression in fathers: A systematic review. Journal of Affective Disorders, 130(3), 358–377. doi: 10.1016/j.jad.2010.06.019 Weinraub, M., Horvath, D. L., & Gringlas, M. B. (2002). Single parenthood. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 109–140). Mahwah, NJ: Erlbaum. Weisman, O., Zagoory-Sharon, O., & Feldman, R. (2012). Intranasal oxytocin administration is reflected in human saliva. Psychoneuroendocrinology, 37(9), 1582–1586. doi: 10.1016/j.psyneuen.2012.02.014 Weisner, T. S., & Gallimore, R. (1977). My brother's keeper: Child and sibling caretaking. Current Anthropology, 18(2), 169–190. Wen, X., Kong, K. L., Eiden, R. D., Sharma, N. N., & Xie, C. (2014). Sociodemographic Differences and Infant Dietary Patterns. Pediatrics, 134(5), e1387–1398. Werner, E. E. (1995). Resilience in development. Current Directions in Psychological Science, 4(3), 81–85. West, M. (1914). Infant care. Care of Children Series, No. 2. Washington, DC: Government Printing Office. West, S. G. (2009). Alternatives to randomized experiments. Current Directions in Psychological Science, 18(5), 299–304.

Whisman, M. A., & Baucom, D. H. (2012). Intimate relationships and psychopathology. Clinical Child and Family Psychology Review, 15(1), 4–13. doi: 10.1007/s10567–011– 0107–2 Whiting, J. W. (1981). Environmental constraints on infant care practices. In R. H. Munroe, R. L. Munroe, & B. B. Whiting (Eds.), Handbook of cross-cultural human development (pp. 155–179). New York, NY: Garland STPM Press. Whitman, T. L., Borkowski, J. G., Keogh, D. A., & Weed, K. (2001). Interwoven lives: Adolescent mothers and their children. Mahwah, NJ: Erlbaum. Wicker, A. W. (1969). Attitudes versus actions: The relationship of verbal and overt behavioral responses to attitude objects. Journal of Social Issues, 25, 41–47. Wilcox, W. B. (1998). Conservative Protestant childrearing: Authoritarian or authoritative? American Sociological Review, 63, 796–809. Wilcox, W. B. (2002). Religion, convention, and paternal involvement. Journal of Marriage and Family, 64(3), 780–792. doi: 10.1111/j.1741–3737.2002.00780.x Williams, L. R., & Steinberg, L. (2011). Reciprocal relations between parenting and adjustment in a sample of juvenile offenders. Child Development, 82(2), 633–645. doi: 10.1111/j.1467–8624.2010.01523.x Williams, P. D., Williams, A. R., Lopez, M., & Tayko, N. P. (2000). Mothers' developmental expectations for young children in the Philippines. International Journal of Nursing Studies, 37(4), 291–301. doi: 10.1016/S0020–7489(00)00004–3 Williford, A. P., Calkins, S. D., & Keane, S. P. (2007). Predicting changes in parenting stress across early childhood: Child and maternal factors. Journal of Abnormal Child Psychology, 35(2), 251–263. doi: 10.1007/s10802–006–9082–3 Wilson, E. O. (1975). Sociobiology: The new synthesis. Cambridge, MA: Harvard University Press. Wilson, P., Rush, R., Hussey, S., Puckering, C., Sim, F., Allely, C. S.,… & Gillberg, C. (2012). How evidence-based is an ‘evidence-based parenting program’? A PRISMA systematic review and meta-analysis of Triple P. BMC Medicine, 10(1), 130. doi: 10.1186/1741–7015– 10–130 Wilson, S., & Durbin, C. E. (2010). Effects of paternal depression on fathers' parenting behaviors: A meta-analytic review. Clinical Psychology Review, 30(2), 167–180. WIN-Gallup International.(2012). Global Index of Religiosity and Atheism. 27 July 2012. Winnicott, D. W. (1975). Through pediatrics to psycho-analysis: Collected papers. Philadelphia, PA: Brunner/Mazel. (Original work published 1948.)

Winstanley, A. V., Sperotto, R. G., Putnick, D. L., Cherian, S., Bornstein, M. H., & Gattis, M. (2015). Consistency of maternal cognitions and principles across the first five months following preterm and term deliveries. Infant Behavior and Development, 37(4), 760–771. Witkin, G. (2012). The modern grandparent's handbook: The ultimate guide to the new rules of grandparenting. New York, NY: Penguin Group. Wittfoth-Schardt, D., Grunding, J., Wittfoth, M., Lanfermann, H., Heinrichs, M., & Domes, G. (2012). Oxytocin modulates neural reactivity to children's faces as a function of social salience. Neuropsychopharmacology, 37(8), 1799–1807. doi: 10.1038/npp.2012.47 Woller, M. J., Sosa, M. E., Chiang, Y., Prudom, S. L., Keelty, P., Moore, J. E., & Ziegler, T. E. (2012). Differential hypothalamic secretion of neurocrines in male common marmosets: Parental experience effects? Journal of Neuroendocrinology, 24(3), 413–421. doi: 10.1111/j.1365–2826.2011.02252.x Woolfson, L. L., & Grant, E. E. (2006). Authoritative parenting and parental stress in parents of pre-school and older children with developmental disabilities. Child: Care, Health and Development, 32(2), 177–184. doi: 10.1111/j.1365–2214.2006.00603.x Yanikkerem, E., Ay, S., & Piro, N. (2013). Planned and unplanned pregnancy: Effects on health practice and depression during pregnancy. Journal of Obstetrics and Gynaecology Research, 39(1), 180–7. doi: 10.1111/j.1447–0756.2012.01958.x Yarosz, D. J., & Barnett, W. S. (2001). Who reads to young children? Identifying predictors of reading of family reading activities. Reading Psychology, 22(1), 67–81. doi: 10.1080/02702710121153 Yarrow, M. R., Scott, P., De Leeuw, L., & Heinig, C. (1962). Child-rearing in families of Working and nonworking mothers. Sociometry, 25, 122–140. Yeager, M., & Yeager, D. (2013). Executive functions & child development. New York, NY:. Norton. Young, K. T. (1991). What parents and experts think about infants. In F. S. Kessel, M. H. Bornstein, & A. J. Sameroff (Eds.), Contemporary constructions of the child (pp. 79–90). Hillsdale, NJ: Erlbaum. Yule, G. U. (1906). On the changes in the marriage-and birth-rates in England and Wales during the past half century; with an inquiry as to their probable causes. Applied Statistics, 69(1), 88– 147. Zahn-Waxler, C., Duggal, S., & Gruber, R. (2002). Parental psychopathology. In M. H. Bornstein (Ed.), Handbook of parenting: Applied parenting (2nd ed., Vol. 4, pp. 295–327). Mahwah, NJ: Erlbaum. Zuckerman, B., & Keder, R. D. (2015). Children in medical settings. In M. H. Bornstein & T.

Leventhal (Vol. Eds.), Ecological settings and processes in developmental systems. In R. M. Lerner (Editor-in-Chief), Handbook of child psychology and developmental science (7th ed., Vol. 4, pp. 574–615). Hoboken, NJ: Wiley Zukow-Goldring, P. (2002). Sibling caregiving. In M. H. Bornstein (Ed.), Handbook of parenting: Status and social conditions of parenting (2nd ed., Vol. 3, pp. 253–286). Mahwah, NJ: Erlbaum.

Chapter 6 Resilience in Development: Progress and Transformation Ann S. Masten and Dante Cicchetti INTRODUCTION HISTORICAL PERSPECTIVES Resilience Science Emerges from Studies of Children at Risk Four Waves of Science on Resilience in Development CONCEPTS AND MODELS Resilience in Developing Systems Defining Resilience Two Judgments The Challenge Criteria: What Are the Risks and Adversities? The Adaptation Criteria: How Well Is the System Doing? Developmental Cascades Adaptation through Time: Pathways of Resilience and Maladaptation Steeling Effects and Posttraumatic Growth Promotive and Protective Processes Biological Models and Processes GxE: Genetic and Epigenetic Models and Processes in Resilience Differential Sensitivity or Susceptibility to Experience and the Environment Biological Models of Protective Parenting A General Framework for Action METHODS Person-Focused Approaches Variable-Focused Approaches Combined Approaches Intervention as Resilience Theory-Testing PSYCHOLOGICAL PROMOTIVE AND PROTECTIVE PROCESSES

Love and Emotional Security Mastery Motivation and Self-Efficacy Intelligence and Problem-Solving Capabilities Self-Regulation Meaning-Making Positive Perspectives on the Self and the Future CHILD-IN-CONTEXT RESILIENCE PROCESSES Family Systems The Special Roles of Schools in Child Resilience Social Networks and Peers Cultural Systems Other Macro-Level Systems and Resilience NEUROBIOLOGICAL PROMOTIVE AND PROTECTIVE PROCESSES Neurobiological Research on Adaptive Systems Implicated in Resilience Molecular Genetic and Neurobiological Processes Epigenetic Studies of Resilience Multiple-Levels in Developmental Reorganization and Plasticity Normative Biological Processes in Resilience Empirical Studies on the Biological Contributors to Resilience Avoiding Biological Reductionism TRANSLATIONAL RESEARCH AND IMPLICATIONS Intervention Research Based on Resilience Models Randomized Controlled Trials Targeting Parent-Child Relationships Randomized Control Trials Targeting Caregiving with Neurobiological-Level Evaluation Moving toward RCTs in Multiple-Levels-of-Analysis Research on Resilience Genetic and Epigenetic Moderation of Intervention Implications of Resilience Science for Policy CONTROVERSIES OLD AND NEW Theoretical Issues

Empirical Issues Is Resilience a Trait? Do Genes Determine Resilience? Is Adversity Good for Children? Is There a Cost to Resilience? PROGRESS AND FUTURE DIRECTIONS Advances in Models and Methods Advances in the Knowledge Base Future Directions REFERENCES

Introduction Resilience science is quintessential developmental psychopathology because it focuses on variation among individuals in relation to their experiences as they adapt and develop. Research on resilience emerged around the same time and in many of the same contexts as the study of developmental psychopathology and this is not a coincidence (Cicchetti, 1990, 2006; Cicchetti & Garmezy, 1993; Masten, 1989, 2007, 2012). Leading scholars concerned with understanding the origins and course of psychopathology and interventions to prevent and treat problems in development began to recognize that adverse experiences played an important role in the lives of many young people who developed problems; however, they also observed that there were striking variations in the adaptive function and life course of children and youth who were believed to be at risk for psychopathology. Influential scholars studying risk for mental illness in human development realized that understanding this variation in adaptive behavior and outcomes was crucial to a full understanding of the origins, prevention, and treatment of problems and disorders (Anthony & Koupernick, 1974; Garmezy, 1985; Garmezy & Rutter, 1983; Rutter, 1979, 1987; Werner & Smith, 1982). The investigators and clinicians who would become pioneers in resilience science recognized that it was vitally important to know why some children do well under very high-risk circumstances while others do poorly in the tasks of life or develop serious mental health problems. They realized, further, that this knowledge could have profound implications for promoting positive development when conditions were not optimal and also for intervening to prevent, ameliorate, or treat symptoms and disorders of psychopathology. The study of resilience in development has had transformational effects on the science and practice of developmental psychopathology. The perspectives and data generated by this work have overturned deficit-oriented models of mental health and intervention, opened up new directions for research, and expanded the range of strategies considered in practices and policies intended to support health and well-being (Cicchetti, 2010, 2013; Masten, 2011, 2013,

2014a, 2014b; Panter-Brick & Leckman, 2013; Prince-Embury & Saklofske, 2013, 2014; Rutter, 2013). The study of resilience has matured and changed over the past half-century of work, in part reflecting the normal progression in any maturing area of inquiry and in part reflecting the dramatic advances in developmental theory and the technologies and methods for research in related fields (Cicchetti, 2010, 2013; Luthar, 2006; Masten, 2014a, 2014c). Resilience science advanced from basic descriptive goals to studies of processes and experiments to test resilience theory by trying to promote positive adaptation in contexts of risk or adversity. Definitions of resilience became more dynamic and multilevel as developmental systems theory became the prevailing perspective in developmental theory and science more broadly considered (Cicchetti, 2010, 2013; Masten, 2011, 2014a, 2014b, 2014c). Advances in techniques for modeling complex data, observing the human brain in action, and studying how genes work and other biological processes provided traction for more sophisticated research on processes proposed in resilience theories or models. There also was growing interest in and efforts to study resilience in more diverse cultural contexts (Ungar, 2012; Ungar, Ghazinour, & Richter, 2013). In this chapter, we highlight the core ideas, findings, and advances in resilience science, emphasizing contributions from the past decade of conceptual and empirical work. Two previous editions of Developmental Psychopathology have included sections on resilience in young people (Luthar, 2006; Masten & Coatsworth, 1995). Our goal in this edition is to provide a current update on the models and findings in resilience science, with a particular focus on the changing definitions and methods emerging from this research domain. Definitions of resilience have shifted over the decades toward a much more dynamic approach, influenced by developmental systems theory. Pathway models and methods have become more prominent. Research on the neurobiology of resilience and cultural processes in resilience are expanding rapidly. This chapter opens with historical perspectives on resilience science, focusing on the goals driving this domain of study. Then we delineate the major concepts and models that define and organize the work in this area, followed by a summary of common methods in resilience research. Next we summarize findings on psychological processes implicated by the evidence on resilience and then we discuss the roles of key socioecological contexts in the resilience of young people, including families, peers, schools, culture, and community. In the following section, we describe the rapidly expanding science on the genetics and neurobiology of resilience. The subsequent section focuses on interventions based on resilience models and the translational implications of the accruing data on resilience for practice and policy, highlighting recent intervention studies designed to test resilience models. In the concluding sections of the chapter, we discuss enduring and new controversies in resilience science and we provide a closing summary of progress and future directions.

Historical Perspectives

Research on resilience in development has many roots. Many of those roots are intertwined with the history of developmental psychopathology and the study of competence in human development (Cicchetti, 1990, 2006, 2010; Masten, 1989, 2014b; Masten & Coatsworth, 1995). These roots include psychoanalytic theory about the protective roles of the ego and mechanisms of defense, the concepts of adaptation and competence in developmental theory, studies of coping and adaptation to stress in clinical psychology and psychiatry, and many other theoretical and empirical efforts to understand variation in the adaptation or development of individuals, particularly those faced with challenges. This history has been discussed in numerous reviews of competence, risk, and resilience over the years (e.g., Cicchetti, 2006, 2010; Cicchetti & Garmezy, 1993; Garmezy, 1981, 1982a, 1985; Luthar, 2006; Masten, 1989, 2012, 2013, 2014b; Masten & Coatsworth, 1995; Masten, Burt, & Coatsworth, 2006; Masten & Garmezy, 1985; Murphy, 1962; Sameroff & Chandler, 1975). The systematic study of variation in adaptation dates to the nineteenth century, shaped particularly by Darwinian and Freudian theory (Masten & Coatsworth, 1995; Masten, Burt, et al., 2006). The “ego” in Freud's theory functioned to facilitate internal and external adaptation amid conflicting needs and goals to meet personal (often unconscious) desires and the expectations or goals of parents and society. How well the personality of an individual managed to balance these competing or conflicting goals would later be described in terms of two dimensions in the personality concepts of Jack and Jeanne Block (1980) as “ego resiliency” and “ego-control.” World War II had a profound influence on the course of psychology and psychiatry, in multiple ways pertinent to the emergence of resilience science (Masten, 2014a, 2014b). The global devastation wrought by the war, with millions of children injured, orphaned, starving, or traumatized by bombing, separations, loss, and many other exigencies of this global conflict, motivated an outpouring of humanitarian aid for children and many efforts to help children affected by the war. These efforts included early research on the effects of traumatic experiences on children as well as intervention to help child survivors of war. Examples of intervention include the establishment of treatment centers for children, such as the Hampstead Nurseries in Great Britain, where Anna Freud and her colleagues treated children affected by the war, and also the founding of the United Nations International Children's Emergency Fund (UNICEF; later broadened to become the UN Children's Fund) in 1946 (Masten, Narayan, Silverman, & Osofsky, 2015). Individuals who would later become leaders in resilience science experienced this war as young people in distinctly different ways (Masten, 2014b). Norman Garmey was a young foot soldier in the European theater of war, fighting at the Battle of the Bulge. After the war, he studied clinical psychology and eventually began studies of risk and premorbid competence that would lead him to observe the variation in adaptive function among children at risk and then on to studies of resilience. Emmy Werner was a young girl in Germany who lived through the bombings, hunger, and destruction resulting from the war in Europe and also experienced the humanitarian aid to children after the war (Werner, 2000). She would go to graduate school in the United States and later lead a classic longitudinal study of resilience, following into adulthood a birth cohort of children born on the Hawaiian island of Kauai in 1955 (Werner &

Smith, 1982). Michael Rutter, a third pioneer in resilience science, was one of the British children evacuated to escape the bombing during the Blitz, a “seavacuee” sent to safety in America for the duration of the war. Subsequently, he returned to Great Britain to study medicine, psychiatry, and epidemiology, conducting extensive research on risk, deprivation, and resilience and publishing many highly influential papers and books, including the most cited article to date on resilience in the psychological literature (Rutter, 1987).

Resilience Science Emerges from Studies of Children at Risk After the war, there was a rapid expansion of research in psychology and psychiatry, some of which was focused on understanding the etiology of mental illness and risk factors in development. There were studies of infants and children at risk due to low birthweight, genetic anomalies, poverty, and other disadvantages (Kopp, 1983; Sameroff & Chandler, 1975) as well as studies of children at risk for mental illness, often because they were related to family members with a known disorder (e.g., Anthony & Koupernik, 1974; Garmezy, 1974; Gottesman & Shields, 1972, 1982; Kopp, 1983; Sameroff & Chandler, 1975; Watt, Anthony, Wynne, & Rolf, 1984). In addition, there were studies of children at risk due to specific adversities, such as death of a parent, divorce, abuse, disasters, and other traumatic experiences (Garmezy & Rutter, 1983). Risk researchers studying children and families soon realized two fundamental facts about the lives of the individuals observed in their research. First, risks and stressors tended to pile up in the lives of many children to the apparent detriment of their adjustment or development. Scores on measures that captured this cumulative risk, including simple tabulations of risk factors or negative life experiences, were associated with less adaptive outcomes. Second, although these risk gradients were often observed, suggesting cumulative effects of adversity, there also was great variation in the adaptive function or development of children with similar levels of risk, suggesting that more was involved than risk alone. The risk researchers who would become pioneering resilience scientists were intrigued by the variation in outcomes among children in their risk studies and particularly by the clear evidence of good adaptation in the presence of high risk for psychopathology or developmental problems. They observed striking cases of individual children who manifested health and competence in the midst of risk and adversity or aftermath of disaster. Those observations and cases motivated the early science on resilience in development (Masten, 2014b).

Four Waves of Science on Resilience in Development Four waves of resilience science ensued (Masten, 2007, 2011; Wright, Masten, & Narayan, 2013).The first wave of resilience science was largely descriptive, as investigators tried to measure the phenomenon of resilience and gain a basic understanding of the qualities of person, relationships, or environment that were associated with better adaptation in conditions of risk or adversity. As investigators gained confidence in the replicated correlates of resilience, they shifted their attention to understanding how protective processes work, and the key question changed from What makes a difference? to How? As hypothesized processes

were studied and confidence grew about key processes, attention turned to identifying and targeting malleable resilience processes through intervention, both in applied clinical contexts and in experiments designed to test hypothesized protective processes, such as effective parenting, mentoring, or self-regulation, discussed later in this chapter. The essential question characterizing the third wave was this: Can resilience be promoted? The fourth wave, now overtaking all the others, emerged from a series of advances in science and technology that revolutionized the study of resilience along with many other aspects of science on development and psychopathology (Cicchetti, 2010, 2013; Masten, 2007, 2014a, b; Wright et al., 2013). These advances included methods for studying the human genome, epigenetics, and the human brain in action as well as statistical strategies for analyzing complex multivariate data. Conceptually, developmental systems theory surged to become the central model of change in human development (Lerner, 2006; Lickliter, 2013; Overton, 2013; Sameroff, 2010; Zelazo, 2013). Concomitantly, dynamic systems theory became a unifying theme in diverse sciences concerned with threats to global well-being and socio-ecological systems, including climate change, disaster, war and terror, economic crises, and computer security, among other issues (Folke, et al., 2010; Gunderson, 2010; Longstaff, 2009; Masten, 2014a; Masten et al., 2015; Norris, Stevens, Pfefferbaum, Wyche, & Pfefferbaum, 2008). In the next sections of this chapter, we highlight concepts and studies that reflect recent advances in resilience theory and methods, particularly those that represent the still-rising fourth wave in the science. Definitions and methods have shifted to emphasize dynamic systems models. A neurobiology of resilience, long anticipated, is finally emerging as methods for studying the role of genes, stress systems, and neuroscience involved in resilience processes improved and became practical. Similarly, there are major advances in research on cultural systems in resilience, also long anticipated. In many ways, the fourth wave engages multiple systems and disciplines in the enterprise of understanding and promoting resilience in human development.

Concepts and Models Concepts central to the study of resilience in development have changed as the study of resilience emerged and matured. The definitions and central ideas in resilience have become more dynamic, and they have aligned with the prevailing models of developmental science more broadly defined. In this section, we examine the central ideas and concepts of resilience science as they have evolved, focusing on the past decade of conceptual advances.

Resilience in Developing Systems Definitions and conceptual models of resilience, basic and applied, have shifted theoretically in concert with a broader shift in human developmental sciences and related fields toward a relational developmental systems framework (Overton, 2013; Zelazo, 2013). This broad perspective integrates theory and ideas from ecological theory (Bronfenbrenner, 1979; Bronfenbrenner & Morris, 1998); general systems theory (von Bertalanffy, 1968);

developmental systems theory (Gottlieb, 2007; Lerner, 2006; Sameroff, 2010); systems thinking in biology (Gottlieb, 2007; Lickliter, 2013); family systems theory (Goldenberg & Goldenberg, 2013; Masten & Monn, 2015); developmental psychopathology (Cicchetti, 2006, 2010, 2013; Gottesman, 1974; Gottesman & Hanson, 2005; Sroufe, Egeland, Carlson, & Collins, 2005); and resilience theory (Cicchetti, 2010, 2013; Masten, 2001, 2013, 2014b). The central themes in a developmental systems perspective on resilience include these eight principles: 1. Human adaptation and development—in both low-and high-risk environments—arise from continuous interactions across levels of function within individuals and between individuals and their environments. 2. Many interacting systems shape the course of development and the processes involved in resilience. 3. The capacity for adaptation can be conceptualized at multiple levels. 4. The capacity for adaptation in challenging circumstances (resilience) depends on many interacting systems. 5. Manifested resilience emerges from many interacting systems and always reflects the current context as well as the history of the child (or system). 6. Living systems are self-organizing with emergent properties that can be surprising or unpredictable from lower levels of analysis. 7. Resilience is dynamic—always changing—because the systems involved in the capacity for adaptation are developing and changing. 8. Resilience—potential or manifested—should not be construed as a trait, although many traits could influence resilience.

Defining Resilience Concomitant with the shift in theory about resilience, definitions have become more dynamic while at the same time becoming more scalable (Masten, 2014a, 2014b). Contemporary definitions emphasize the processes of adaptation and the distributed capacity for adaption rather than traitlike notions of inherent “hardiness” or the “ability” within a system to bounce back from difficulties. Efforts to address mass-trauma events, such as natural disasters, war, terror, and epidemics, as well as concerns about global climate change, have spurred interest in diverse fields to consider resilience processes that involve multiple systems at different levels of analysis (e.g., Gunderson, 2010; Masten, 2014a, 2014b; Masten et al., 2015). Given goals of integrating theory and findings coming from different fields of inquiry, important for understanding and responding effectively to global challenges, interest grew in communicating findings across fields and integrating knowledge across levels of analysis. These goals motivated efforts to define resilience in ways that would work across levels and disciplines that typically may

focus on one or two specific levels (e.g., resilience in neural systems, individuals, families, communities, ecologies, computer systems, etc.). Scalability of terminology can be very helpful for communication and integration. A small group of investigators that included the first author worked on this issue of terminology and scalability in a network on resilience funded by the National Science Foundation (see Longstaff, 2009; Masten & Obradović, 2008). This group presented their ideas at Resilience 2008, the first of several ongoing international conferences on resilience sponsored by the Resilience Alliance, Swedish Academy of Sciences (the first meeting was held in Stockholm), and others. Subsequently, Masten (2011, 2014a, 2014b) adopted a systems-oriented definition of resilience that would be workable for other disciplines focused on diverse kinds and levels of systems. Given the advantages of a scalable definition of resilience to describe the processes and capacity for adaptation of any dynamic complex system in the context of serious threats or adversities, we think it is useful to define resilience for scalability across levels of dynamic, developing systems. Resilience broadly refers to the potential or manifested capacity of a dynamic system to adapt successfully through multiple processes to disturbances that threaten the function, survival, or development of the system. This definition can be applied to many dynamic, developing systems, including an individual child, a family, a community, an emergency response system, an economy, or various ecological systems in the context of threats to the global climate. For the purposes of this chapter, which is focused on the resilience of human individuals, we define resilience in this way: Resilience is the potential or manifested capacity of an individual to adapt successfully through multiple processes to challenges that threaten the function, survival, or positive development. Many systems are involved in the processes leading to successful adaptation in a child, family, or community. Moreover, systems that are interconnected across levels will influence the resilience of each other. In other words, the resilience of an individual child that is manifested and observable at the level of behavior depends on the operation and interactions of many other systems, both within the child (immune system, stress response system, etc.), in relationships or family resilience, or in the larger sociocultural and ecological systems in which that child's life and development are embedded.

Two Judgments Identifying resilience in the individual case or for the purposes of research requires two criteria, as noted for some time by resilience scholars (e.g., Garmezy, 1974; Garmezy, Masten, & Tellegen, 1984; Luthar, 1991; Luthar, Cicchetti, & Becker, 2000; Masten & Coatsworth, 1998). Resilience refers broadly to positive adaptation in a context of risk or adversity. Thus, two constructs must be considered in making inferences about resilience, whether one is simply thinking of an example of a single case of manifested resilience or one is operationalizing the core constructs in order to conduct a large-scale research project on resilience. The “risk/adversity” criteria and the “adaptation” criteria have been defined in

multiple ways in the literature over the years, reflecting the complexity of the issues but making it difficult to synthesize the findings, particularly through empirical strategies such as metaanalysis. Many critiques of the resilience literature have noted this problem (Luthar, 2006; Masten, 1999, 2012, 2014b). Nonetheless, there is broad agreement that resilience refers to a subset of adaptive processes or outcomes when there is significant threat. In situations where there is little or no adversity experienced, we would describe children in terms of how well they are doing or developing rather than their resilience. The construct of resilience assumes that there is exposure or potential exposure to serious challenges or adversities to the function or development of the system under consideration. In the next section, we discuss the two major components for defining resilience in more detail.

The Challenge Criteria: What Are the Risks and Adversities? Concepts and measures of risks or various threats to the well-being or development of a child have been considered for decades in the developmental sciences (Evans, Li, & Sepanski Whipple, 2013; Garmezy, 1974; Kopp, 1983; Obradović, Shaffer, & Masten, 2012; Sameroff & Seifer, 1983). As noted, the study of resilience emerged from the study of risk as investigators noted the variability in adaptive patterns among children considered to be at risk for psychopathology. A wide range of risk factors—predictors of specific problems or generally undesirable developmental outcomes—were identified, including premature birth, abusive parenting, exposure to trauma in war or disaster, and the disadvantages of poverty, among many others. Risk researchers quickly realized that the factors associated with elevated probabilities of mental health symptoms, academic problems, health issues, and other undesirable outcomes rarely appeared in isolation and commonly piled up in the lives of the same children. This realization led the risk scholars to focus on cumulative risk and its effects on development (Masten, Best, & Garmezy, 1990; Rutter, 1979; Sameroff, Seifer, & Bartko, 1997). Many of the cumulative risk concepts and measures were modeled on the study of risk in medicine. Sameroff (2006) has described the history of composite risk scores dating back to the Framingham heart study, where it was observed that a cumulative total number of risk factors, rather than any one risk in particular, predicted heart problems. Numerous studies of cumulative risk emerged, with many investigators observing risk gradients, where problem or symptom levels tended to risk as risk levels rose (Evans et al., 2013; Obradović et al., 2012). SES gradients linked health outcomes to cumulative risk scores based on socioeconomic disadvantage, with much higher mortality and morbidity rates found among those with low socioeconomic status (SES; Adler & Ostrove, 2006). In studies of war and disaster, these gradients were often termed dose gradients when it was found that trauma symptoms, for example, were often higher with greater exposure to danger, loss, destruction, and other indices of life-threatening intensity of adverse experiences (Masten & Narayan, 2012). The risk researchers also recognized the complexity of the processes involved in stressful life experiences, such as divorce, death of a parent, homelessness, extreme poverty, and other

adversities. Divorce, for example, was recognized as a process punctuated with many possible adverse experiences along with a number of potentially positive experiences (Hetherington, 1979; Kelly & Emery, 2003). Two major types of measures of cumulative risk became prominent: tallying life events (weighted or unweighted by objective or subjective magnitude scales) and counting up established risk factors into a composite risk score. Rutter (1979) and Sameroff and colleagues (Sameroff, 2006; Sameroff, Seifer, Barocas, Zax, & Greenspan, 1987) were highly influential in modeling strategies for compositing risk factors. The other main strategy utilized life event questionnaires or interviews to measure the level of adverse life experiences that a person or family had experienced in a specified time window (e.g., past year, lifetime). The study and assessment of negative life events by questionnaire was pioneered by Holmes and Rahe (1967) and also Brown and Harris (1978), who developed a widely-used method of structured, contextual life interviews. These methods were adapted for studies of children by Coddington (1972a, 1972b), the Project Competence research group (Masten & Tellegen, 2012), and others that wanted to assess exposure to stressors in the form of heterogeneous life experiences (e.g., Compas, 1987; Goodyer, Kolvin, & Gatzanis, 1985). Hybrid methods also developed, combining methods to create more complex indices of cumulative adversity in a child's life, utilizing a combination of lifetime life event questionnaires, life history charts, and similar methods (e.g., Gest, Reed, & Masten, 1999). A recent variation on the life event questionnaire method is the Adverse Childhood Experiences (ACEs) measures, typically a set of 10 or 11 items completed by adults about their childhood history. This approach was used initially to survey adults in California (Felitti et al., 1998); subsequently it has been used by various states in the United States to depict the rates of exposure to major risk factors in childhood (as recollected by teens or adults), such as child maltreatment or parental incarceration, and link this to adolescent or adult health and mental health indicators (e.g., Minnesota Department of Health, 2011). Research with ACEs indicates striking risk gradients where poor health outcomes are related to the number of ACEs. ACEs and other clearly negative life experiences have negative effects when they occur but may have little or no effects on development when they do not occur. However, many of the “risk factors” associated with problems in adjustment are actually continuous variables that have a positive pole, such as good versus poor parenting. Resilience scholars have pointed out that these measures of risk are simultaneously measures of “assets” or resources because they reflect bipolar dimensions (Masten, 2001; Masten & Reed, 2002). The same measures can be flipped over to produce “asset gradients” where the probability of desirable outcomes rises as a function of cumulative assets. The Search Institute has capitalized on this reality to create a widely used measure of 40 “developmental assets” that can be used to gauge the support of communities for young people; this group recognized that it was powerful and motivating to show communities the positive side of the story—that youth with more of these assets were doing well (Lerner & Benson, 2002; Scales, Benson, Leffert, & Blyth, 2000; Scales, Benson,

Roehikpartain, Sesma, & Van Dulmen, 2006; Sesma, Mannes, & Scales, 2013). Their approach underscored the importance of adding assets rather than reducing risks. However, whether one focuses on risk gradients or asset gradients, many of the cumulative risk indices obscure the variation within a particular level of risk, high or low. Pioneers in the study of resilience brought attention to the off-gradient cases that were not well predicted by high cumulative risk scores (Masten, 2014b). Some individuals were doing much better than might be expected from their level of risk. Historically earlier diathesis–stress models of psychopathology had proposed in a similar way that some individuals were more vulnerable to adversity, doing worse than might be expected by the level of risk (Gottesman, 1974). Offgradient cases in either direction (worse or better than expected) suggested that there might be moderating influences on risk or adversity that could not be captured by additive effects. These interaction effects became the focus of research on protective factors that might explain betterthan-typical or expected outcomes as well as the diathesis studies of genetic vulnerability and other moderating influences that could lead to worse effects. Cumulative risk scores also could mask the salient effect of particular experiences on wellbeing or function. These scores were designed to package risk because risk factors often cooccurred, and the number of risk factors often seemed to be more important than the nature of the particular risk exposures. However, there are situations where this may not be the case, particularly in situations of extreme trauma. In a meta-analysis of posttraumatic stress (PTS) studies, Furr, Comer, Edmunds, and Kendall (2010) were able to show that in disasters (masstrauma events including natural disasters and terror attacks), greater PTS symptoms were found in disasters with greater loss of life and loss of loved one and friends. In a specific analysis to unpack risk in their study of Bosnian adolescents exposed to war and political conflicts, Layne and colleagues (2010) identified particular experiences as more potent risk factors, including exposure to life-threatening experiences and traumatic death. In their study of former child soldiers, Betancourt et al. (2010) described the “toxic” effects of rape and killing others on long-term well-being and adjustment. Many of the cumulative risk models and gradients were conceptualized and tested using linear models. However, there also was consideration of nonlinear models, at least in theory (Garmezy et al., 1984; Masten et al., 1988). The “challenge” model was curvilinear, where the relation of adaptive function to stress or adversity was quadratic—with the best outcomes at mid-levels of stress rather than very low or very high levels of challenge, akin to the familiar inverted “U” function of the Yerkes-Dodson law relating performance to arousal. The concept of possible steeling effects of some adversity exposure also was nonlinear (Rutter, 1987). In recent years, nonlinear models of risk have become more prevalent as the statistical strategies and data collection needed for testing these effects became more widespread and feasible to execute. Intriguing data from studies of children in war and political conflict have suggested nonlinear dose effects (see Masten et al., 2015). For example, some research on child soldiers exposed to extraordinary trauma in captivity suggest that there may be asymptotic effects where the exposure is so high that everyone has passed the threshold for full expression of posttraumatic symptoms; at this level of risk, symptoms are less related to

exposure dose of traumatic experiences and more related to the quality of the recovery environment (Klasen et al., 2010). Another study of youth engaged in political conflict, Palestinian youth in Gaza, by Qouta, Punamäki, and El Sarraj (2008) found nonlinear relations of engagement to conflict exposure. At lower to middle levels of exposure to the conflict, they observed a typical dose-response effect, with adaptive behavior declining as exposure levels rose. However, at extremely high levels of exposure to extreme political violence, youth appeared to rally for the cause. Another important theme that became salient in recent studies of adversity, particularly in regard to mass-trauma adversities, was the role of media exposure in risk (Masten & Narayan, 2012; Masten et al., 2015). Evidence has accumulated that exposure to adversity through media coverage can have significant dose effects on children, and particularly young children. Moreover, studies that examine the potential risk of social media exposure are just getting under way.

The Adaptation Criteria: How Well Is the System Doing? Resilience science focused on the processes that make it possible for some individuals to weather the storms of adversity. What are the criteria by which resistance, recovery, or posttraumatic growth are judged? Here, too, many criteria have been the focus of study, some broad and some narrow, some external and observable, others internal and subjective. To judge or study resilience, one must evaluate how well a system is doing by some kind of criteria. In research on human resilience, positive adaptation, adjustment, or outcomes have been judged by different criteria at different levels of analysis (Cicchetti, 2010; Luthar, 2006; Luthar et al., 2000; Masten, 2007, 2014b; Masten, Burt, et al., 2006; Schoon, 2006). With the study of resilience emerging from research on children at risk for major mental disorders or antisocial behavior, sometimes the criteria were defined in negative ways, by avoiding psychopathology or manifesting few symptoms (e.g., avoiding substance abuse or conduct disorder). A study by Nigg and colleagues (2007) provides an example; they examined protective factors associated with avoiding the development of attention-deficit/hyperactivity disorder, oppositional defiant disorder, or conduct disorder in children at risk due to family adversity. However, some investigators focused on how well children were doing in the developmental tasks of childhood, adolescence, or early adulthood on multiple dimensions of competence, discussed further in the next paragraph (e.g., Masten & Tellegen, 2012). These measures of adaptation or adjustment were often behavioral (e.g., academic achievement, getting along with peers) and age-related, changing as children matured and entered new contexts. Some investigators focused on subjective well-being or happiness as key criteria for adapting well (e.g., Jordan & Graham, 2012). Studies of immigrant youth and acculturation stress often focus on both psychological well-being and adjustment in school (Motti-Stefanidi, 2014). Developmental studies, particularly longitudinal studies, often measured adaptation by achievements in the tasks that are commonly expected for individuals to accomplish at different ages (McCormick, Kuo, & Masten, 2011). For example, success in age-salient developmental

tasks were key criteria for the assessment of adaptation in three influential longitudinal studies of risk and resilience: the Children of Kauai (Werner & Smith, 1982, 1992, 2001), the Minnesota Longitudinal Study of Risk and Adaptation (Sroufe, Egeland, Carlson, & Collins, 2005), and the Project Competence Longitudinal Study (Masten & Tellegen, 2012). In early childhood, these developmental tasks include forming attachment bonds with caregivers, learning to walk and talk, and beginning to comply with rules. Older children are expected to go to school, learn to read and write and calculate with numbers; get along with peers and teachers; and follow the cultural rules and laws of their societies. Adolescents and young adults are often expected to form deeper friendships, develop romantic relationships, form their own families, and enter the workforce, higher education, or take on child-rearing responsibilities. Although it was recognized long ago that adaptation criteria are grounded in cultural and historical context (Masten & Coatsworth, 1995, 1998), empirical attention to cultural variations in the criteria for adaptive behavior or competence gained considerable traction in the past decade of resilience science (Masten, 2014a; Panter-Brick & Leckman, 2013; Ungar, Ghazinour, & Richter, 2013). Work of the Resilience Centre in Halifax, Nova Scotia, provides examples of the collaborative international efforts under way to study cultural similarities and differences in defining good adjustment, as well as protective processes, in resilience.

Developmental Cascades Over the past decade, there also has been a surge of empirical studies focused on developmental cascades across domains and levels of adaptive function. Developmental cascades refer to progressive or spreading changes across systems, time, and even generations, resulting from multiple-level interactions in complex living systems, that have cumulative effects on the course of a system's development (Masten & Cicchetti, 2010c). These effects were noted long ago in research on the etiology and course of psychopathology and also in the literature on the development of competence (Burt, Coatsworth, & Masten, Chapter 9, this volume; Cicchetti & Cannon, 1999; Cicchetti & Tucker, 1994; Cole & Maxwell, 2003; Dodge, Greenberg, Malone, & the Conduct Problems Prevention Research Group, 2008; Hanson & Gottesman, 2007; Hinshaw, 1992; Hinshaw & Anderson, 1996; Kagan, 2005; Kohlberg, LaCrosse, & Ricks, 1972; Masten & Coatsworth, 1998; Masten et al., 2005; Patterson, Reid, & Dishion, 1992; Rutter, 1999; Rutter, Kim-Cohen, & Maughan, 2006; Rutter & Sroufe, 2000). These effects have been called snowball effects, amplification effects, positive chain reactions, and spillover effects among other terms for similar concepts. Basically, cascades describe the cumulative consequences of the interactions among systems that shape the course of development, recognizing that changes spread in dynamic, interdependent systems. Developmental systems theory forecasted and empirical findings supported the systems idea that problems or successes in one domain or level or part of interacting systems (e.g., within a person and a family or a peer group) could spread to affect other domains, levels, or people (see Masten & Cicchetti, 2010a, 2010b). Research on resilience highlighted the cascade effects pertinent to positive adaptation, as well as promotive and protective processes. One central theme focused on the theory and findings

that competence begets competence as children develop. Achievements in core developmental tasks in one period of development set the stage for future competence. This idea has been championed not only by developmental psychologists (Masten & Coatsworth, 1998; Sroufe, Egeland, Carlson, & Collins, 2005), but also by economists and prevention scientists concerned with benefit-to-cost analysis (Heckman, 2006; Temple & Reynolds, 2007). Investing in competence early in development was recognized as a highly cost-effective strategy for both promoting later competence and preventing psychopathology, because of cascading consequences in development. A second central theme focused on interrupting negative cascades or intervening to prevent negative cascades from unfolding through strategic prevention or intervention efforts (Burt et al., Chapter 9, this volume; Masten & Cicchetti, 2010c). Interventions can be designed to target mediating processes in anticipated developmental cascades sequences, timed and targeted to prevent negative and promote positive consequences of system interactions. Basic research on naturally occurring resilience highlighted the potential for cascading consequences of deliberate efforts to reduce dose and mitigate negative consequences of adversity and stress on development. Examples of interventions to interrupt or generate cascades are discussed further in subsequent sections.

Adaptation through Time: Pathways of Resilience and Maladaptation Resilience theory also highlighted positive pathways of adaptation through time. Pathway models of adaptive function have a long history, dating back to the work of early embryologists describing development. Examples of early pathway models include the epigenetic landscape of Waddington (1957/2014) and pathways delineated by Gottesman and his collaborators (Gottesman, 1974; Gottesman & Shields, 1982) portraying the combined effects of epigenetic processes in shaping the life course toward and away from schizophrenia. In resilience science, theoretical pathway models represented efforts to capture the course of adaptive function over time, in the relation to changes in adversity and other circumstances. In the literature on resilience in children, Masten and Reed (2002) delineated a number of pathways that had been described in both theory and individual case histories of resilience. These included pathways of stress resistance, breakdown and recovery, and normalization. Later, Masten and colleagues would differentiate pathways for acute challenges, such as those observed in disasters, terrorist attacks, and accidents, from those commonly observed in chronic adversity exposures, such as family dysfunction with domestic violence or substance abuse, entrenched poverty, and child neglect or abuse in the family (e.g., Masten, 2013; Masten, Monn, & Supkoff, 2011; Masten & Obradović, 2008). For example, Masten and Narayan (2012) illustrated a number of theoretical pathways often described in the literature on children in war and disaster. Bonanno (Bonanno, 2004; Bonanno & Diminich, 2013) have delineated similar pathway models in the literature focused on adult resilience. A sample of pathway models are illustrated in Figures 6.1 and 6.2 showing patterns of different pathways in the context of acute trauma or a period of chronic adversity followed by

a period with improved context (due to natural changes in adversity or interventions to reduce adversity). In the aftermath of acute trauma, these manifested resilience patterns have been observed and theoretically proposed: stress-resistance, where functioning remains about the same despite the challenges; breakdown and recovery, where there is initial decline in function followed by return to more or less the original level of function; and posttraumatic growth, where function improves in the aftermath of a severe challenge, sometimes following an initial decline, as if the individual's functioning is galvanized by the demands of responding to the challenge and/or the processes of recovery. During the period of severe, chronic trauma, children do poorly or decline in function, as illustrated during the first period of time in Figure 6.2. Then, when conditions improve for child function or development, as capacity for resilience builds or replenishes, resilience pathways in the form of recovery or normalization become evident.

Figure 6.1 Resilience Pathways After Acute Trauma. Model of resilience pathways after acute trauma. The gray zone represents normal or typical adaptive function. Path A represents stress resistance. Path B illustrates breakdown and recovery. Path C represents posttraumatic growth. Source: © Ann S. Masten. Reprinted with permission.

Figure 6.2 Resilience Pathways after Chronic Adversity. Model of resilience pathways showing a period of chronic exposure to severe adversity followed by a period when conditions improve. Path A shows declining adaptive function during chronic adversity followed by recovery when conditions improve. Path B shows a normalization pattern when conditions improve. Source: © Ann S. Masten. Reprinted with permission.

With the advent of advanced statistical modeling and more longitudinal studies using repeated measures, it has become feasible to examine and test such models empirically. Readily available methods of growth mixture modeling or group-based trajectory modeling developed by Nagin, Muthén, and their colleagues have made it possible to identify pathways followed by groups of individuals within a longitudinal sample (Jung & Wickrama, 2008; Muthén & Muthén, 2000; Nagin, 1999, 2005; Nagin & Odgers, 2010; Nagin & Tremblay, 1999). Although some of the resulting studies emphasized pathways of symptoms or problems rather than competence or positive adaptation (see Nagin & Odgers, 2010, for examples), these analyses frequently identified relatively healthy or positive trajectories too (paths with stable low levels of symptoms or trajectories with declining symptoms). In fact, studies of symptom pathways following disasters reveal that the majority of children or youth show improving or stable good functioning following acute disasters (e.g., La Greca, Lai, Joormann, Auslander, & Short, 2013). Betancourt and colleagues also observed that in a majority of cases, trajectory analyses indicated resilience in war-exposed youth followed over time from Sierra Leone, many of them former child soldiers with extremely high and chronic exposure to trauma (Betancourt, McBain, Newnham, & Brennan, 2013). Their results for internalizing symptoms indicated four trajectories, with 41% of young people showing stable, low symptom levels and

an additional 48% showing an improving trajectory over time. Only 5% showed stable, highlevel symptoms, and 6% showed deteriorating patterns. Although the advent of mixture/trajectory modeling has facilitated empirical studies of resilience pathways, it is interesting to note that there appear to be more examples modeling symptoms than positive criteria of adjustment. There are very few examples of mixture modeling to identify resilience trajectories based on positive criteria (e.g., Obradović, Burt, & Masten, 2006) even though there have been some attempts to identify trajectories of academic achievement in normative samples. Davis-Kean and Jager (2014), for example, used growth mixture modeling to analyze data from the Early Childhood Longitudinal Study-Kindergarten Cohort, a nationally representative sample. They examined achievement trajectories within racial groups, which revealed achievement gaps and interesting “catch-up” trajectories within racial categories.

Steeling Effects and Posttraumatic Growth One of the most intriguing pathways reported anecdotally in resilience case studies is posttraumatic growth, when adaptive function appears to improve following exposure to adversity (particularly acute catastrophes). Some have suggested that there is an initial breakdown of function followed by a reorganization or strengthening process (Calhoun & Tedeschi, 2006). Some families report better function after dealing with a crisis or a child's illness. Colville and Cream (2009), for example, found that many parents of children who were in a pediatric intensive care unit reported posttraumatic growth after the children left the hospital. On closer inspection, they also noted that this effect was curvilinear—it was more commonly reported by parents with moderate levels of PTS than those with higher or lower perceived stress levels. The idea that some exposure to adverse experiences and challenges is better than none or too much exposure for longer-term adaptive function also is a concept with a long history in resilience science (Masten, 2012). Rutter (1987) described the steeling effects of exposure long ago in his classic article on resilience, and he continues to garner the evidence on these nonlinear effects (Rutter, 2012). The vaccination model of prevention is built on an exposure model (to a weaker or less-threatening form of the threat; Masten, 2014b). Adaptive systems often require calibration or experience to “learn” to work properly or tune to the expected environment; this is true of the immune system and our muscles, and may well be true of behavioral adaptation. Such findings have led to a flurry of articles on the sometimescontroversial theme that Nietzsche (2013) so succinctly described in 1888: “What does not kill me makes me stronger.” The controversy comes with the idea that trauma is “good for you,” which is very problematic in situations of child maltreatment and other devastating adversities. Practice with adaptation may be helpful but only in modest amounts that do no harm. Surviving Ebola may provide immunity, but the risk of dying and cost to health and well-being are astronomically high. That said, the possibility that adapting to adversity can alter the organism in positive ways for future adaptation is fundamental to preventive interventions, including stress-inoculation

training as well as immunizations against childhood diseases. An overprotected child may not end up with the adaptive capacity for responding successfully to the vicissitudes of life that will inevitably come. Nonlinear effects of early life adversities also have been studied in animal models and particularly in primates (Lyons & Parker, 2007; Parker & Maestripieri, 2011). Here too evidence has accumulated that the relation of adversity experiences to later adaptive function is quadratic rather than linear. Loss, abuse, or deprivation of caregiving can have calamitous effects on the development of primates, similar to human infants. However, some experience with brief separations in infant macaques and other primates apparently can result in better regulated stress-responses systems that yield greater resilience later in life. This effect in squirrel monkeys has been described as a stress-inoculation effect (reviewed by Parker & Maestripieri, 2011). Some experience with manageable separations, in other words, can be better for long-term adaptation than no experience with separations.

Promotive and Protective Processes The search for factors that could account for the variation in short-or long-term outcomes among children exposed to adversity resulted in the identification of variables or conditions that seemed to make a difference. However, one of the first issues that came up in this search was distinguishing generally good predictors of positive outcomes (in various domains of function or development) regardless of risk from predictors that appeared to play a particularly important role when risk or adversity was high. Eventually these two kinds of effects came to be called “promotive” factors (a term popularized by Sameroff, 2000) and “protective” factors. In essence, this distinction differentiated effects observed at most or all levels of risk from those effects observed in conditions of challenge. Moreover, it was not the factors that truly mattered, but instead their function in varying conditions, because it was quickly observed that the same factor could function both ways. Parenting quality and good cognitive skills have been extensively implicated as both correlates/predictors of achievement in developmental tasks and also as moderators of risk or adversity (Masten, 2014b; Masten & Coatsworth, 1998). Pioneering investigators often commented on this issue. Garmezy et al. (1984), among others, pointed out that this distinction from an analytical perspective reflected “main effects” in contrast to “interaction effects.” Rutter (1987) was an early champion of keeping these functional roles distinct, and he has continued to argue (Rutter, 2012) that studies of resilience and protective processes related to resilience should be focused on what happens when adversity is high. Promotive factors (which also can be called promoters, assets, and resources) are positive correlates and predictors of desirable outcomes. Some of these may be purely positive, but in many cases, the correlates of positive development or adjustment are actually dimensional attributes (e.g., good versus poor parenting) that could easily be “reversed” to become a risk measure, as already noted in the discussion of risk.

Similarly, moderators of risk or adversity that predict variations in adaptive behavior or other outcomes in the context of threats to adaptive function could be described as “protective” or “vulnerability” indicators in terms of their function. But in the situation of high challenge or threat to adaptive function, the “expected” norm references help to define the presumed nature of the factor's role or function. Thus, for example, if the expected typical response to increasing adversity or threat is a decline in function, and the moderating action appears to be at the high end of the purposed moderator, the variable will be described as a protective factor. If one is low on the moderator and the outcome is worse than generally expected or observed, and the action is concentrated at the other end of the dimension, that effect will be described as a vulnerability effect. And both effects could be observed. The normative levels of the moderator and adversity/threat in the general population also influence the interpretation of a moderator as a protective factor or a vulnerability. If average levels of parenting warmth or structure appear to protect children from the effects of high adversity exposure, then parenting effects will be viewed as protective. Poor parenting, in contrast, could exacerbate the consequences of even modest exposure to adversity, often interpreted as a vulnerability effect. The essential idea in defining these effects is not the factor is protective or promotive per se but instead that it functions in a particular way in a given context or situation with respect to a particular criterion of outcome. This idea was emphasized in early reviews (e.g., Masten, Best, & Garmezy, 1990; Rutter, 1987), but it has been underscored by recent theory and evidence on differential sensitivity to experience, discussed further in the section on “Differential Sensitivity or Susceptibility to Experience and the Environment.” In a negative environment, rife with adversity, “sensitivity” could be viewed as a vulnerability, whereas in an enriched and safe environment, or in the context of an intervention, it could be enhancing to development. The functional significance of “sensitivity” varies in different conditions.

Biological Models and Processes Early resilience investigators were strongly influenced by the biological models prevailing at the time, particularly in medicine and public health, including psychiatry and epidemiology (Masten, 2014b). These models included numerous concepts and models already discussed, such as cumulative risk, diathesis-stressor models of disease, and vaccination, that all explained how multiple influences might account for the development of problems or illnesses and, concomitantly, how to think about prevention. The ideas of stress-inoculation and steeling effects drew on models of vaccination, where exposure to a weaker or safer form of a dangerous biological agent could stimulate the immune system to make antibodies that subsequently would protect the organism from the worst ravages of dangerous diseases like smallpox or polio. Similarly, challenging the muscles by exertion would initially break down function but, with recovery, lead to strengthening of those muscles in the future. In these models, clearly it is not “no exposure” that is best but rather some kind of manageable exposure to a challenge that is followed by recovery and yields a more adaptive organism. Modern examples include the hygiene hypothesis for asthma, where lower exposure to

pathogens or microorganisms in development can lead to higher risk for asthma in the future. Explanations for the hygiene effect are complex (see Figueiredo et al., 2013), but the general idea is that modern lives may be “too clean” to tune or program the individual's immune system effectively. Growing up on a farm with exposure to a range of microorganisms can be viewed as protective for some kinds of allergies (Guerra & Martinez, 2008; Okada, Kuhn, Feillet, & Bach, 2010; von Mutius & Radon, 2008). In these models, developmental timing is important because protection is conferred by early exposure that helps to calibrate the immune system whereas later exposure to the same environmental exposures can result in allergic reactions. There also is a growing body of evidence based on animal models that repeated exposures to moderately stressful experiences can have positive effects on how young animals handle future stressors (Parker & Maestripieri, 2011). Contemporary variations of these biologically-based challenge models include child-rearing or educational ideas where parents or teachers aim to build psychological strength by exposing children to manageable challenges. Another example can be found in interventions based on stress-inoculation training that are designed to prepare children for potentially traumatic experiences, such as hospitalization or terror attacks. In Israel, for example, school children have been trained to handle attacks related to political conflict and terror, with natural experiments showing that such training can be effective for reducing the risk of PTS (e.g., Wolmer, Hamiel, & Laor, 2011). It is noteworthy that the models and empirical evidence on the effects of challenging experiences, whether bio-logical or psychological, suggest that it is modest levels of challenge, with adequate resources and supports for recovery, that may help to prepare a child or a primate for future challenges. There is little evidence that overwhelming trauma and adversity make children stronger. Another early biological model was provided by sickle-cell anemia, where having two alleles for this genetic disorder caused a life-threatening illness yet having one allele could confer some protection from malaria in certain regions where the disease was endemic. One of the resilience pioneers, Michael Rutter (1987), noted this model in a classic article on resilience. This biological model indicated the importance of context for interpreting risk and protective effects, both the genetic context and the rearing environment. This example illustrated the phenomenon of a balanced polymorphism, where the heterozygous individual has a survival advantage but only under some conditions. Over the past decade, advances in the development of increasingly accurate and practical tools to measure genes and epigenetic processes have revolutionized the science in multiple fields, including medicine and the study of adaptive function and development more broadly defined. These advances have influenced research in developmental psychopathology (Grigorenko & Cicchetti, 2012) and resilience (Kim-Cohen & Gold, 2009; Kim-Cohen & Turkewitz, 2012). Consequently, new models of resilience and resilience-based intervention have emerged.

GxE: Genetic and Epigenetic Models and Processes in Resilience Gene–environment interaction (GxE), one of several distinct forms of gene–environment

interplay (Rutter, Moffit, & Caspi, 2006), refers to observable variations in individual functions or development that are attributable to the interdependence between a specific identified variation in the DNA sequence of a gene and a well-defined and carefully measured experience or feature of the environment, often a pathogen (Caspi, Hariri, Holmes, Uher, & Moffitt, 2010; Moffit, Caspi, & Rutter, 2006). In GxE, environmental experiences moderate genetic effects (or vice-versa) on adaptive function and development, including normal, pathological, or optimal outcomes (Cicchetti & Rogosch, 2012; Rutter, 2006). For example, genetic effects on outcomes may be observed only in certain environmental contexts or in conjunction with different histories of experience; conversely, experience may relate only to outcomes among individuals with specific genetic characteristics. Risk-exposed individuals who carry the “protective” version (or allele) of a gene with respect to that risk condition develop significantly lower levels of psychopathology compared to comparably risk-exposed individuals with the “vulnerable” allele (Kim-Cohen & Gold, 2009). Although there has been an abundance of research conducted on GxE interactions and evidence for GxE interactions in the etiological pathways of psychopathology and manifested resilience (Caspi et al., 2010; Karg et al., 2011), it is important to note that a number of criticisms have been levied against GxE research (see, e.g., Duncan & Keller, 2011; Duncan, Pollastri, & Smoller, 2014; Keller, 2014). These include small sample sizes that can yield low statistical power, failure to control for all potential confounders in all models and failure to compute all relevant interaction terms in the same model, and a paucity of replication studies (Duncan & Keller, 2011; Keller, 2014). Advances in molecular biology have helped to usher in a renascence of interest in the contribution that studies on GxE can make to unraveling the complex pathways of adaptive and maladaptive development, including resilience processes (Cicchetti & Blender, 2006; Grigorenko & Cicchetti, 2012a, 2012b; Rutter, 2006). These advances include the completion of the DNA sequencing of the human genome (Collins, Morgan, & Patrinos, 2003; Cowan, Kopnisky, & Hyman, 2002) and the publication of the map of human haplotypes that provides valuable information about individual genetic variation and a powerful tool for identifying both vulnerability and protective genes (D. C. Crawford & Nickerson, 2005; Insel & Quiron, 2005). As Rutter et al. (2006) articulated, an understanding of the complexities involved in comprehending normal, abnormal, and resilient patterns of development may prove useful in avoiding misleading types of stigma and biological reductionism (see also Curtis & Cicchetti, 2003; Hinshaw & Cicchetti, 2000). Moreover, comprehending the complex variation in dynamic multilevel pathways emphasizes the importance of genes in all developmental trajectories involving risk and protection. The empirical contributions of a molecular genetic approach enable us to discover genetic elements that contribute to developmental processes of normality, psychopathology, and resilience without requiring foreknowledge of the underlying biochemical abnormalities (Ciaranello et al., 1995). These contributions make the search for the intermediate developmental mechanisms in the gene-environment-brain-behavior interconnections more accessible than ever before (Gottesman & Gould, 2003; Gottesman & Hanson, 2005; Hanson & Gottesman, 2007; Lenzenweger, 2013; Moffit et al., 2006). Moreover, progress in molecular

genetics also raises hope of developing interventions to prevent and remediate mental disorder and to promote resilience (Belsky & van IJZendoorn, 2015; Cicchetti & Blender, 2006; Cicchetti & Curtis, 2006; Grigorenko & Cicchetti, 2012a, 2012b; Howe, Beach, & Brody, 2010; Masten, 2014b).

Differential Sensitivity or Susceptibility to Experience and the Environment One of the most influential set of ideas in GxE theory over the past decade with implications for resilience science stems from the concepts of differential susceptibility (Belsky, Bakermans-Kranenburg, & van IJzendoorn, 2007; Belsky & Pluess, 2009) and biological sensitivity to context (Boyce & Ellis, 2005; Obradović & Boyce, 2009). According to the evolutionary-derived hypothesis of differential susceptibility to the environment, individuals vary in their developmental plasticity or malleability to negative or positive experiences (Belsky & Pluess, 2009). Genes that confer risk in harsh environments are thought to confer benefits in normal or nurturing environments (Belsky et al., 2007; Belsky & Pluess, 2009; Ellis, Boyce, Belsky, Bakermans-Kranenburg, & van IJzendoorn, 2011). In other words, the characteristics of individuals (including their gentoypes) that render them disproportionately more vulnerable to effects of experiencing adversity may also make them disproportionately more likely to benefit from supportive contexts (Belsky et al., 2007; Boyce & Ellis, 2005; Ellis et al., 2011). Differential susceptibility in the view of Belsky and colleagues (2007) can be distinguished from related ideas, including diathesis-stressor theory (Gottesman & Shields, 1972), by evidence on effects in negative and positive environments. If susceptibility is evident only in negative environments, but not in positive environmental contexts, then “diathesis” is indicated. In contrast, differential susceptibility theory is supported by a cross-over interaction such that susceptible individuals are disproportionately affected “for better and for worse” by both positive and negative experiences (Belsky et al., 2007). Because a simple phenotype cannot be adapted to all potential conditions, the creation of variability within families increases the probability that at least some offspring will be successful in every generation (Belsky & Pluess, 2009). Therefore, differential susceptibility to the environment presumes that individuals vary in their susceptibility to context as a result of genetic differences. Differential susceptibility to the environment also provides a possible explanation for individual variability in responsiveness to intervention (Ellis et al., 2011). This variation is particularly important because most treatment outcome studies, even those yielding statistically significant effects and having medium to large effect sizes, include many individuals who do not benefit from the intervention (Belsky & van IJzendoorn, 2015). Thus, identifying differential susceptibility markers is a key to tailoring treatments to uniquely fit individual differences. As such markers are identified, they may be examined as potential moderators of resilience or of intervention outcomes. Theoretically, while there may be “sensitive” children who are responsive to both bad and

good environments, there also could be traits or other endogenous factors in a child that yield receptivity only to positive environments. This possibility led to a further differentiation of individual differences to describe the sensitivity to positive environments or advantaged conditions, described as “vantage sensitivity” by Pluess and Belsky (2013), along with Manuck (2011). Substantially less research has been directed toward investigating endogenous factors that are associated with variation in response to positive influences. Vantage sensitivity could be viewed as the bright side of differential susceptibility theory (Pluess & Belsky, 2013). Vantage sensitivity is used to describe the notion that some persons are more sensitive and positively responsive to the environmental advantages to which they are exposed (Pluess & Belsky, 2013). Vantage sensitivity also can be differentiated from a general sensitivity to good or bad environments. Some individual differences in developmental responsiveness to environmental experience may emerge only under positive or supportive conditions. In these instances, vantage sensitivity, not differential susceptibility—which is characterized by disproportionate sensitivity to both positive and negative exposures and experiences—would be indicated (Pluess & Belsky, 2013).

Biological Models of Protective Parenting Many other models of resilience could be drawn from the biological literature, including animal models. A complete treatment of this topic is beyond the scope of this chapter. Just one additional model example will be highlighted, which is the rat model of parenting developed by Michael Meaney and colleagues that dramatically illustrates the power of environments to moderate genetic risk through epigenetic pathways (Karatsoreos & McEwen, 2013; Meaney, 2010). Animal models have played a key role in understanding how both risk and protective influences may work. Meany and colleagues demonstrated through a series of elegant experiments that maternal behavior in rat mothers, even foster mothers, can alter gene expression, with lasting effects on the adaptive behavior and vulnerability to stressors among the pups. Additional studies in animals and humans support the idea that qualities of parenting can alter the effective genotype and thereby influence later vulnerability or resilience in the face of adversity (Karatosreos & McEwen, 2013). Here again, developmental timing matters. The high-versus low-licking behavior of rat mothers has its greatest impact during a sensitive period early in pup development.

A General Framework for Action As resilience models and corresponding research strategies developed, a parallel resilience framework for action also emerged, because children in urgent situations cannot wait for a complete science to guide translation of knowledge into practice (Masten, 2011, 2014b). Action often may be required before the evidence we ideally would want to have is in place. In these situations, action can be guided by general principles based on the science to date. Resilience perspectives on programs and policies usually emphasize these principles: Set positive goals.

Include positive influences and outcomes in models of change as well as risks. Monitor and measure positive achievements and progress as well as problems. Increase promotive influences and mobilize adaptive systems while also reducing harmful exposure to risk or adversity. Consider multiple levels and systems for leveraging change. Intervene strategically during windows of opportunity for promoting positive cascades and interrupting negative cascades. These principles do not ignore or neglect risk, recognizing that preventing adversity or toxic risk exposures often yields a high return on investment, both financially and socially. Additionally, these principles underscore the importance of timing, for preventing risk as well as for promoting positive processes: interventions are more effective when they are well-timed in terms of developmental plasticity or contextual transitions during periods when developing and interacting systems may be more amenable to change (Masten, 2014b).

Methods Resilience research required methods to explore and test hypotheses about adaptation in the context of risk or adversity. These methods included strategies for assessing the primary components of resilience models—including risk, adaptation, and potential mediators or moderators discussed in the section on models—but also methods for studying patterns and processes in the lives of people who did or may manifest resilience. In this section, we discuss two major traditions for studying resilience in children, person-focused and variable-focused in approach, as well as recent hybrid methods that combine features of both and intervention studies as a method for studying resilience.

Person-Focused Approaches Single cases of manifested resilience have played a prominent role in the history of resilience science and continue to play key roles in inspiring new lines of work and in communication of research findings. Scientists often report that their interest in resilience was spurred by case examples of individuals in their own or other studies, particularly the pioneers in psychology and psychiatry who propagated the first wave of research (Masten, 2014b). These pioneers set out to study the origins of psychopathology among children at risk for various reasons but often were struck by the compelling individual examples they observed of young people who were doing well despite high-risk status or exposure to severe adversity. Garmezy (1982b) wrote an article titled “The Case for the Single Case” in which he described the heuristic value of individual case examples for providing insights, inspiration, and clues for future research. He often described individual cases when he taught, spoke, and wrote about resilience, including famous examples (e.g., Queen Elizabeth I) and anonymous examples from newspaper accounts or his studies. There are published case studies of resilience, although they are far fewer than the

biographical accounts available in books, films, and other media. One example is the case of “Sara” (Masten & O'Connor, 1989), an account of a high-risk child (of a mentally ill mother) who experienced traumatic family losses around 1 year of age followed by an inadequate foster case placement with strangers. Growth failure and developmental decline in multiple areas of function followed until she was hospitalized for a full evaluation at the age of 30 months. On recommendation of the treatment team, Sara was referred for adoption and the “treatment” of an experienced, stable, loving family. Sara was followed over time, and she responded to her permanent placement with a dramatic developmental rebound, returning to a normative developmental trajectory in physical, cognitive, and social development. A blind-tohistory follow-up assessment when Sara was 6 years showed a continued course of positive development. In their book on the High Valley Resilience Study, Hauser, Allen, and Golden (2006) tell the stories of nine individuals from a longitudinal study of 67 young people hospitalized for emotional and behavioral problems during adolescence who managed to “get out of the woods” as they grew up. The authors focused more deeply in this narrative on four of these resilient youth and how they overcame a troubled adolescence, identifying three major protective factors often observed in larger studies of resilience: positive relationships, agency, and reflective self-control. By moving beyond the single case and comparing a subgroup of cases that showed resilience by their criteria, Hauser et al. illustrate the aggregated case method, which has a number of advantages over the single-case approach. The key drawback to the single case study is generalizability. It is not clear to what degree a single life represents unique processes versus broad, generalizable findings. Although a single case can be highly compelling and heuristic, investigators are motivated to turn to the power of numbers to make a more convincing case about their hypotheses. Nonetheless, even in large studies of resilience, scientists often single out single case examples to illustrate and communicate their findings. Aggregated case studies focus on the whole person, as do single case studies, but aim to identify larger samples of persons who show resilience as defined for the particular study. One classic approach begins with identification of a high-risk group known from past evidence to have a higher-than-normal probability of some undesirable outcome due to a well-established risk factor, high cumulative risk, or adversity exposure. Then a subgroup of individuals who are doing well is identified within the high-risk group. The High Valley Resilience Study followed this approach. Some of best-known data from the influential resilience study of a birth cohort born on Kauai by Werner and Smith (1982) also followed this approach. The investigators identified a high-risk subgroup based on biological and environmental risk factors present before the children were 2 years old (e.g., poverty, family discord, perinatal stress, mental or physical illness in a parent). Then they examined how well members of the high-risk group were doing in subsequent follow-ups on multiple criteria of adjustment. These criteria at ages 10 and 18, for example, included indicators of conduct, achievement, and social competence. Werner and Smith identified a “resilient” subgroup of 72 youth who were then compared to less successful peers from the high-risk category at different points in development in order to identify individual and family differences associated with resilience.

Focusing only on high-risk youth has some limitations. Without low-risk comparison groups, for example, it is not possible to distinguish promotive and protective factors. A difference may be found on cognitive skills, personality, or parenting quality, but it will not be clear if the effect of the factor is any different than it would be among a low-risk sample. Factors that show similar roles under low and high risk—promotive factors—have “main effects” from a statistical point of view, showing the same relation to the adaptive criterion of interest at any level of risk. In contrast, protective effects suggest a more important role when risk is high. In other words, the two groups may differ on an individual or family variable of interest in exactly the same way for a high-risk sample as they would in a low-risk sample, which suggests a correlate of adjustment but does not suggest a special role when risk is high. The extended classic approach adds one or more low-risk comparison groups, so that promotive and protective effects can be differentiated. The Project Competence Longitudinal Study (Masten & Tellegen, 2012) utilized this approach in some person-oriented analyses, cross-classifying participants into groups by multiple criteria of competence as well as cumulative levels of adversity. Either cluster analysis or cut-scores can be used to identify the groups. This strategy led to interesting findings on similarities and differences among the different subgroups as well as a nearly “empty cell” suggesting that there were few young people in a school-based sample with a history of low risk who were maladaptive. Results indicated many similarities in the two groups showing positive adjustment around age 20, lowrisk and high-risk, and many differences between these two groups of adaptive youth and their high-risk/maladaptive peers. In some situations, there is no comparable “low-risk” group; virtually everyone has been exposed to the risk or adversity, although there may be subgroups with lower exposure. This can happen when a population is exposed to mass-trauma events, such as hurricanes, terror, war, or disease pandemic. Studies of survivors of the Khmer Rouge genocide in Cambodia encountered this problem (Masten, 2014b). Sometimes the low-risk comparison group differs in so many ways from the high-risk group that it is difficult to control for all the ways the two groups differ other than the key difference of interest. An example might be comparing children who grow up in foster care, who often have a long and complex history of risk, to children raised in stable family situations who have lower risk but also may differ in many aspects of sociodemographic status. A contemporary twist on these classic person-focused strategies is trajectory analysis. These strategies, discussed later in this section, are designed to extract subgroups from longitudinal patterns of adaptation among high-risk populations based on intraindividual change over time in a variable of interest, and thus they share features of person-focused and variable-focused methods (Muthén & Muthén, 2000).

Variable-Focused Approaches Person-focused methods consider the lives and attributes of individual people as a whole organism, often comparing similarities and differences in people identified on the basis of history and current function as showing resilience. Variable-focused methods are designed to

study patterns of variance and different ways that variables are linked, drawing on the power of statistical methods to identify potential clues to processes that might explain resilience. Variable-focused methods often provide more nuanced clues to key processes involved in resilience. Common models of variable effects often investigated in resilience studies are shown in Figures 6.3 to 6.7. Figures 6.3 and 6.4 illustrate several ways of showing the roles of multiple risks (negative) or assets/promotive factors (positive) or bipolar variables in predicting a desirable or positive outcome. From a statistics perspective, both figures represent “main effects” models.

Figure 6.3 Risk or Promoter Gradient. A cumulative risk or promoter gradient illustrating how adaptive function falls as accumulating risk factors rise or promotive factors fall. Stars represent individuals who are functioning well even though their risk levels are high. Source: © Ann S. Masten. Reprinted with permission.

Figure 6.4 Risk and Promoter Model. Basic “main effects” model of risk and promotive factors that influence an adaptive outcome of interest. This model includes an intervention with a promoter effect. Source: © Ann S. Masten. Reprinted with permission.

Figure 6.3 represents a typical risk or dose gradient where adaptation is better at lower levels of risk and/or higher levels of promoters (assets or resources). In Figure 6.3, at the high-risk end of the x-axis (which may also indicate low levels of promotive factors or assets), adaptation is typically poor (symptoms are high or achievement is low or both). However, the stars represent two cases where an individual is doing much better than might be expected for the person's level of risk. These stars specify off-gradient cases that may signify resilience. Individuals may adapt at levels better than expected due to moderators operating to mitigate risk or unmeasured promotive influences that are in play, discussed further later. Another way to depict these additive models with risks, bipolar predictors, or promoters is shown in Figure 6.4. This simple model includes one bipolar variable (risk/asset), one risk factor, one naturally occurring promotive factor, and an intervention intended to act as a promoter that was added to a child's life (represented by the dashed line), such as providing a tutor. Mediating effects are illustrated in Figure 6.5. Here the effect of a risk factor (e.g., economic stress or maltreatment) on the adaptive outcome (e.g., child function or achievements; child competence) is mediated by an intervening variable, such as parenting quality or executive function skills. Mediating models set forth a hypothesis of how risks (or interventions to mitigate risk effects) work. An intervention (shown as a dashed line) has targeted the mediator as a strategy for changing the consequences of adversity or risk on child outcome. This intervention strategy illustrates a potential moderator effect (discussed later) on a key mediator of risk. For example, interventions have targeted parenting quality in families or executive function skills in children in an effort to mitigate the effects of disadvantage on school success or psychopathology (e.g., Zelazo & Carlson, 2012; Masten et al, 2014; Patterson et al., 2010). These types of interventions can focus on mitigating risk effects on the mediator or boosting support for the mediator.

Figure 6.5 Mediated Risk Model with Intervention. Risk affects an intermediate variable that in turn changes adaptive outcome. This model includes an intervention intended to indirectly improve the outcome by mitigating risk effects on the mediator or by supporting the mediator. Source: © Ann S. Masten. Reprinted with permission.

Moderator models are shown in Figure 6.6. There are two primary forms of moderators: independent moderators of risk, such as personality factors (e.g., emotional reactivity) that enhance or reduce sensitivity to a particular kind of risk or stressor, and risk-dependent moderators, where the risk or stressor triggers or activates the moderator. The independent moderator might include individual differences in personality or biology that influence how an individual usually responds to stress (e.g., stress-reactive or calm). The risk-activated moderator is analogous to an airbag in a motor vehicle or the immune response to an infection. Emergency services for children and families are examples of risk-activated protective factors, intended to mitigate harm to children or families by providing emergency services or supports. Interventions in this model take the form of boosting the risk-independent protective influences and adding or mobilizing more or better “emergency services.” Parents and teachers, for example, can learn to respond more effectively to emergencies and threats of various kinds that threaten children in their care. Stress-inoculation strategies of prevention seek to prepare children, families, or schools for what to do in emergencies in an effort to boost risk-activated protective processes.

Figure 6.6 Moderated Risk Model. Moderated risk model showing one independent moderator that mitigates or enhances the effects of the risk (adversity) variable and one risk-activated moderator that is triggered by risk or adversity and then buffers or exacerbates the effects of the risk/adversity. Source: Adapted from “Ordinary Magic: Resilience processes in development,” by Ann S. Masten, 2001, American Psychologist, 56, pp. 227-238.

Figure 6.7 illustrates two additional variations on moderating influences stemming from the theories on differential susceptibility or biological sensitivity to context already discussed. Differential sensitivity moderates both positive and negative experiences, enhancing their effects, which is good in a favorable environment and bad in an unfavorable environment. Vantage sensitivity moderates only favorable (advantaged) environments. Not shown in this figure is the third kind of moderator that enhances only negative experiences, which would be described as vulnerability.

Figure 6.7 Differential Sensitivity Model. Moderator model of differential sensitivity to experience showing one “differential sensitivity” moderator that enhances or attenuates the impact of both bad (adversity) or good (advantage) experiences and a second “vantage sensitivity” moderator that moderates the impact only of good (advantage) experiences. Vulnerability (a moderator of adverse experiences) is not shown. Source: © Ann S. Masten. Reprinted with permission.

Combined Approaches There are two approaches to analyses of resilience that combine features of person-centered and variable-centered analyses. Some investigators conduct multiple sets of analyses to examine the data from different perspectives. In the Project Competence Longitudinal Study, for example, investigators combined methods focused on variables, including structural equation modeling and regression analyses, with classification of people into groups of high or low risk and adaptation, a classic person-focused approach (Masten, 2014b; Masten & Tellegen, 2012). This group also has presented case material from the study, carefully disguising the identity of the participants (Masten, Morison, Pellegrini, & Tellegen, 1990). In the Kauai study, Werner and Smith (1992, 2001) also reported multiple methods, some focused on person and others focused on variables. Approaching longitudinal data sets from multiple methodological perspectives can be illuminating. A different strategy that combines both approaches is methods that examine distinct trajectory patterns of intraindividual growth and change, using the advanced statistical techniques already described, often called group-based trajectory modeling (Nagin, 1999, 2005) or growth mixture modeling (Muthén & Muthén, 2000). These methods are used to identify distinct

groups of people who follow a similar trajectory or pathway of functioning on a variable of interest over time. Using techniques pioneered by Nagin, Muthén, and others (see Jung & Wickrama, 2008; Nagin & Odgers, 2010), groups of children or youth showing different patterns of function over time can be extracted from longitudinal data with repeated measures. These groups can then be compared on other variables that might explain the different patterns of adjustment. As noted in the discussion of pathway models, these mixed-model approaches have been used to identify subgroups of children and adolescents who show different recovery patterns after trauma experiences, including hurricanes (La Greca et al., 2013) and prolonged trauma as child soldiers (Betancourt, McBain, Newnham, & Brennan, 2013). In both these examples, the majority of individuals showed improving or good function over time, consistent with theoretical models of resilience pathways. Mixed models also can be used to demonstrate treatment effects in interventions studies and to identify subgroups responding better (or worse) to treatment (Nagin & Odgers, 2010). Through propensity score matching, comparison groups can be studied of people on the same trajectory, some of whom were randomly assigned to receive an intervention. These tools are valuable for research to develop and test interventions based on resilience theory.

Intervention as Resilience Theory-Testing One of the most powerful methods for testing theories and hypotheses about resilience is experiments with random assignment to treatment, testing a well-delineated model of change. Randomized controlled trials are the gold standard of evidence on what works to promote change and also for establishing causal evidence. Children cannot be assigned to random adversity or risk exposure, of course, but ideas about promotive and protective processes can be tested by intervention designs where the goal is to promote resilience in children or youth. Numerous intervention studies have attempted to promote better outcomes in young people at risk by mitigating risk, boosting resources, or improving protection in order to foster resilience (Masten, 2014b). Targets of these efforts include: the reduction of premature birth and all its consequences by prenatal care (Alexander & Kotelchuck, 2001); the reduction of insecure and disorganized attachment in maltreated youngsters through improving the quality of parent-child attachment or interaction (e.g., Cicchetti, Rogosch, & Toth, 2006; Dozier, Peloso, Lewis, Laurenceau, & Levine, 2008;); promoting school engagement (Anderson, Christenson, Sinclair, & Lehr, 2004; Hawkins, Kosterman, Catalano, Hill, & Abbott, 2005); providing after-school programming (Durlak, Weissberg, & Pachan, 2010; Lauer et al., 2006; Mahoney, Larson, & Eccles, 2005); building self-regulation skills (Diamond & Lee, 2011; Raver, 2012); and improving parenting (Patterson, Forgatch, & DeGarmo, 2010; Sandler, Schoenfelder, Wolchik, & MacKinnon, 2011). There are numerous other examples, both of interventions designed to test resilience theory and of interventions designed with other goals in mind that provide compelling evidence that it is possible to promote resilience. In addition, resilience theory and models suggest that change can be initiated at different system levels or multiple levels simultaneously (Masten, 2014b). Change cascades across systems, so that an intervention targeting the parent can spread throughout the family or an

intervention targeting child behavior change also changes the parents, siblings, or family as a whole (Masten & Monn, 2015). Changes at a biochemical level in the brain, perhaps from medication or experience-induced epigenetic changes, can cascade upward into behavior, thereby alleviating depressed mood and leading to changes in parent behavior toward children or broader family changes (see Masten & Cicchetti, 2010). The challenge for inducing change based on resilience science is knowing enough to optimally target the who, what, and when of leverage and opportunity for change: who to target, at what system level, and when, in terms of developmental timing or a cascade progression. Although prevention science is beginning to address this challenge, evidence to date is sparse but growing.

Psychological Promotive and Protective Processes Five decades of research on resilience consistently implicate a set of promotive and protective factors across wide-ranging situations and cultures, suggesting that there are fundamental adaptive systems that promote and/or protect human development and function under challenging circumstances (Masten, 2001, 2007, 2014b). Very early reviews and books (e.g., Garmezy, 1985; Garmezy & Rutter, 1983; Rutter, 1979, 1987) as well as more recent reviews and volumes on the empirical literature (e.g., Goldstein & Brooks, 2013; Luthar, 2006; Masten, 2014b; Rutter, 2013) have identified these commonly observed correlates or predictors of good adaptation among children at risk due to various adversities or risk. From a developmental systems perspective, the processes involved in these correlates and predictors of resilience would be multiple level and complex, spanning interactions from the cellular and genetic to the socioecological level. In this section we review these key adaptive systems at the level of psychosocial function, the level of focus when they were initially studied. The neurobiological processes of resilience related to these adaptive systems, which is a focus of intense interest in contemporary research, are discussed subsequently in the chapter.

Love and Emotional Security All reviews on what matters for human resilience have pointed to the importance of relationships across the life span (Luthar, 2006; Masten, 2014b; Rowe & Kahn, 1998). The nature of the relationships varies across development, with the quality of caregiving and the attachment bonds with caregivers very salient in early development. As children grow older, relationships with other adults and friends or romantic partners become important for resilience. With respect to resilience, the theorized and observed functions of any of these relationships are multifaceted and multidimensional. In developmental theory, attachment theorists have proposed that beginning early in development, humans (as well as other social mammals) evolved to form powerful and special bonds with their caregivers and other individuals involved in their lives (Ainsworth, 1989; Bretherton & Munholland, 1999; Cassidy & Shaver, 2008; Sroufe, 1979; Sroufe et al., 1999; Thompson, 2000). These relationships were naturally selected because of their survival and

reproductive value. The human attachment system with primary and other important caregivers organizes in the latter part of the first year of life, providing emotional security, safety, and also motivation for learning. Between caregiver and the young child, attachment is bidirectional. In the event of a threat perceived by either party, the attached individual will seek proximity or other means to connect and ensure felt security. A frightened child will crawl, walk, and eventually run to the caregiver; a frightened parent, fearful for a child's safety, will also take action to ensure the child is safe or soothe the child. In the disaster literature, the physical proximity of attachment figures reduces stress and trauma symptoms (Masten & Narayan, 2012; Masten et al., 2015). The attachment system functions as a nurturing and safety system, but it also is strongly related to learning and exploration. A young child who feels secure is more likely to explore, play, and learn and often takes observable joy in sharing accomplishments and interacting with the attachment figure. The attachment system is linked with the mastery motivation system, discussed in the next section. Responsive and sensitive caregiving predicts secure attachment relationships that are then carried forward into relationships with teachers, friends, or romantic partners that play helpful roles later in a person's life (Sroufe et al., 2005; Thompson, 2000). At the same time, highquality parenting of this kind serves the immediate function of co-regulation so that children who do not yet have independent self-regulation skills to modulate arousal, emotion, or behavior have an “external” regulator at hand. Thus, good parenting also would be expected to relate to the development of self-regulation, another adaptive system implicated in research on resilience. Parents and other caregivers serve many other functions as promoters and protectors of development (Masten 2014b; Sroufe et al., 2005). Parents function as teachers of language, life skills, and culture, and they also protect children from immediate and anticipated dangers, seek medical attention for them, and soothe their anxieties, among many other functions. Given these many roles for development in normative and dangerous situations, a capable parent/caregiver is central to healthy development and resilience. Not surprisingly, the greatest threats to the early well-being and development of a child occur when the primary caregiver is lost, disabled, or unable to serve these roles, such as when a parent is killed or neglectful. In disasters or war, when children lose or are separated from caregivers, restoring contact or functions of this system is viewed as an emergency (Masten & Narayan, 2012). Even worse is the situation when the caregiver her- or himself poses threats to the child, when a parent is the source of maltreatment or danger. This situation can pose double jeopardy to a child who faces adversity without the protection of a fundamental adaptive system in the form of an attached and capable caregiver. Nonetheless, even in the context of maltreatment, at least one good relationship with a caregiver shows protective effects on child function and development (Alink, Cicchetti, Kim, & Rogosch, 2009; Collishaw et al., 2007; Egeland, Carlson, & Sroufe, 1993; McGloin & Widom, 2001). Given the salience of secure attachment relationships and the effective parenting that often accompanies these relationships, it is not surprising to find that interventions for high-risk

children focused on restoring, improving, or otherwise bolstering the functions and quality of parenting have been successful. Interventions to improve parenting and foster care, for example, have shown efficacy in controlled experiments for promoting better outcomes in children at risk for behavior and emotional problems (e.g., Patterson et al., 2010; Sandler, Schoenfelder, Wolchik, & MacKinnon, 2011; Toth & Manley, 2011).

Mastery Motivation and Self-Efficacy A powerful reward system for adaptation and learning, often described as “mastery motivation,” has been linked to both competence and resilience in development (Masten, 2014b; Masten et al., 1995). People get feelings of satisfaction and pleasure from perceptions of their own agency and accomplishments, evident from an early age. Toddlers newly able to walk or to run take great delight in what they can do. Robert White described this motivation for agency as “effectance” and experience of satisfaction that accompanied perceived effectiveness in the environment as “efficacy.” White (1959) argued in a classic paper on this powerful biological system was naturally selected because of its value for adapting successfully. Other psychologists and developmental scientists have studied the apparent motivation for learning and adapting in psychological concepts of “self-efficacy” (Bandura, 1997) or “intrinsic motivation” (Ryan & Deci, 2000). Individual children and adults who score high on mastery motivation measures or self-efficacy tend to respond to challenges or failures by persisting or trying again rather than giving up (Bandura, 1997). Hauser, Allen, and Golden (2006), in their book on teens who succeeded in the years following psychiatric hospitalization, noted the importance of “agency and the quest for mastery” in the youth showing a resilient pathway. Laub and Sampson (2003) similarly were struck in their follow-up research by the role of agency and satisfaction in the lives of the men from the famous longitudinal study of delinquency by Sheldon and Eleanor Glueck (1950) who were able to desist from a life of crime. Parents and teachers often reinforce this system by providing opportunities for children to experience pleasure in their own accomplishments (Masten, 2014b). They break down learning and other challenges into manageable steps and encourage children to try on their own so that children get a sense of accomplishment and develop a can-do attitude toward challenges. This system has been relatively neglected compared to some of the other prominent adaptive systems associated with resilience, yet it seems likely that many of the intervention programs that promote competence in parents or children are engaging this system and aligning with the natural reinforcement and pleasure that humans gain from attempting and succeeding in something challenging but doable.

Intelligence and Problem-Solving Capabilities Since the earliest research on resilience, the capacity for problem-solving by children (or adults) also has been identified as a key promotive and protective factor for adaptive function and development, in both low- and high-risk situations (Luthar, 2006; Masten, Burt, et al., 2006). Intelligence tests were developed to predict adaptive success, particularly in a school

context, so this finding also was not surprising (Masten, 2014b). Classic studies of resilience over the years have found that cognitive skills or assessed intellectual abilities forecast better adaptation among children from high-risk backgrounds, particularly in regard to better grades and fewer behavior problems (e.g., Loësel & Farrington, 2012; Masten & Tellegen, 2012; Werner & Smith, 1992, 2001). Good intellectual function appears to afford protection for children at risk for conduct problems, possibly through multiple pathways, including school success or engagement, better planning or self-regulation skills, and greater likelihood of affiliating with prosocial rather than deviant peers. Additional pathways arise from correlated promotive or protective factors that often accompany good cognitive functioning, including competent parents and socioeconomic advantages; children with good cognitive skills are more likely to have competent families, live in resource rich and safer neighborhoods, and attend better schools. Highly aversive environments for child development may impair cognitive development through multiple pathways, beginning prenatally. What some investigators describe as “toxic stress” can undermine brain development and the potential of a child for resilience (Shonkoff, 2011). Better cognitive skills may well reflect better nutrition, health care, parenting, and opportunities to learn, all potentially mediated by healthier brain development and the development of specific cognitive and learning skills, including various aspects of memory and cognitive control. Good cognitive function and development in a context of very high risk may be an indication that multiple aspects of child development have been protected despite adverse childhood experiences. Recent research on cognitive development and its role in resilience has focused particularly on executive function and other self-regulation skills, discussed in the next section, as well as the role of parent-child interaction and quality of early childhood education for the development of good cognitive and self-regulation skills and school readiness (Masten, 2014b). Interventions to promote competence and resilience in young children have targeted executive functions and early literacy or numeracy (Blair & Raver, 2012; Diamond & Lee, 2011; Zelazo & Carlson, 2012).

Self-Regulation Resilience studies often report that more adaptive children and youth from high-risk or adverse backgrounds have better self-regulation skills compared to less successful peers (Zelazo & Carlson, 2012; Masten, 2014b). Self-regulation of attention, emotion, cognition, and action is associated both with current and future competence in achieving developmental tasks and also in responding effectively to adversity (S. M. Carlson, Zelazo, & Faja, 2013; Duckworth, 2011; Masten & Coatsworth, 1998; Rothbart, 2011). Research on self-regulation is burgeoning, although the ideas linking self-control to adaptive behavior date back more than a century to concepts of the ego in psychoanalytic theory (Masten, 2014b). The multifaceted job of the ego in this theory was to regulate internal experiences of anxiety and pleasure while ensuring that external adaptation to the environment (including the rule for social conduct) was appropriate. Over the past decade, research on executive function (EF) skills has expanded rapidly,

including basic investigations and intervention research (Zelazo & Carlson, 2012). What were originally developed as laboratory tests and methods to assess or boost EF skills have moved into classrooms, homes, and even emergency shelters for homeless families in research aimed at understanding nature and roles of these skills in high-risk populations (Masten et al., 2014; Zelazo & Carlson, 2012). Interventions to promote EF skills through teacher, parent, or child training show promise as strategies for promoting resilience among high-risk children.

Meaning-Making Human capacity for imbuing life with meaning has been implicated in autobiographies on resilience and also in empirical studies. Man's Search for Meaning, Viktor Frankl's (1959/2006) account of survival in a concentration camp during World War II, provides one of the most compelling accounts of human capacity for finding beauty and meaning in the midst of great suffering. Generations of readers, including school children the world over, have been moved by Anne Frank's diary, another classic example of human ability to construct meaning in a context of great adversity and the comfort that such beliefs can hold. The act of “telling the story” or constructing a narrative may be an important form of meaning-making in action for survivors of extraordinary trauma. A recent example is provided by Elizabeth Smart's 2013 book about her life in captivity after she was kidnapped, where she tells her story of anguish, faith, and survival (Smart & Stewart, 2013). In the book, Smart described how her faith in her family's love and her religion sustained her hope in the face of ongoing imprisonment, rape, and hunger. Religions, and associated spiritual faith of individuals following a religious system, reflect shared systems of beliefs about the meaning of life, among many other protective practices and beliefs offered by cultural systems handed down across generations (E. Crawford, Wright, & Masten, 2006). Older youth and adults in resilience studies report that their spiritual faith and beliefs that life has purpose sustained them during period of great adversity or that a religious conversion marked a turning point in their lives (e.g., Elder & Conger, 2000; Werner & Smith, 1992, 2001). Although religious beliefs and practices can apparently serve a protective role, religion also can lead to extremism and participation in violence (E. Crawford et al., 2006). Similarly, gang leaders and war leaders in political conflicts can provide meaning for life and may recruit young people, particularly adolescents, by exploiting the need for meaning in a context of despair and conflict (Masten, 2014b). Meaning-making, like other fundamental human adaptive processes, can be directed toward goals that other people would not view as evidence of resilience but rather as destructive or immoral. Findings in research on voluntary youth participation in political violence (see Barber, 2009, 2013) indicate that marginalized young people, otherwise hopeless or oppressed, may find a sense of purpose and feelings of mastery through their active participation in political conflicts, including acts of violence. One of the most provocative issues in contemporary resilience science, as well as for societies in search of peace, is the co-option of such adaptive systems for antisocial or violent purposes. The concept of meaning making is complex, and the empirical literature on processes that may

be involved lags behind the conceptual literature (Park, 2010). Moreover, little data on young people is available. In the theoretical literature focused on adults, a person's beliefs, goals, subjective sense of purpose, and motivation have been conceptualized as key elements of a global sense of meaning (Park, 2010). Even in the adult literature, it is not clear whether preexisting meaning, searching for meaning, or finding meaning in relation to a traumatic experience is important and how (Bonanno, Westphal, & Mancini, 2011). One of the rare domains of research focused on meaning making in youth has investigated meaning in relation to identity development. McLean and Pratt (2006), for example, studied identity development and turning points in the transition years from adolescence into early adulthood. Their study spanned ages 17 to 23, including measures of identity status and optimism along with writing a narrative about important transitions or changes in life, called a “turning point narrative.” These narratives were analyzed for meaning making through a qualitative coding strategy, with the authors concluding that their results corroborated other evidence that “the redemptive story sequence is associated with meaning The management of difficult life events by using meaning-making processes to find the good within the bad is important to the development of a healthy narrative identity” (p. 721). Redemptive meaning refers to finding meaning in suffering or adversity. McLean and Pratt also found that general optimism was associated with more mature narratives on meaning. Narrative identity is a research domain focused on the processes and adaptive significance of observations that people construct narratives or stories about their lives that provide a sense of coherence and identity (McAdams & McLean, 2013). A salient theme in these narratives is what McAdams, McLean, and their colleagues refer to as “redemptive sequences,” where a negative experience leads to some kind of positive outcome. Investigators of narrative identity often code constructs such as agency and meaning making in their qualitative analysis of narratives that have played a salient role in accounts and research on resilience. McAdams and McLean (2013) also noted the influence of culture on the personal narratives of young people. Positive mood or subjective well-being has been linked to perceived meaning in life in the studies of narrative identity in youth (McAdams & McLean, 2013) as well as in adult research on meaning in life (King, Hicks, Krull, & Del Gaiso, 2006). The direction of the causal influence is not clear yet. In numerous accounts of resilience, beliefs that life has meaning are strongly associated with positive perspectives on the self and future. Research on narratives about the self includes studies of neural processes that may be involved in constructing these narratives and also intervention studies. In a paper on brains creating stories of selves, D'Argembeau and colleagues (2014) used functional magnetic resonance imaging (fMRI) to study neural processes during tasks of autobiographical remembering. Their results are consistent with the idea that there are individual differences in which aspects of neural function are engaged during personal narratives, with differences in self-reflectiveness playing a role. Their findings suggest possible neural processes linked to meaning making in the context of autobiographical memory. Finally, it is interesting to note that some evidence-based interventions for youth who have experienced trauma combine a sharing of narrative with desensitization techniques. Narrative

exposure therapy (NET) and the child version of this method (KIDNET) involve symbolic and narrative accounts of the life story combined with exposure techniques of desensitization. For example, using a rope to represent the life course, rocks (bad experiences) and flowers (good experiences) are laid out and gradually discussed. Narrative exposure therapy has been effective in controlled trials with recovering child soldiers (Ertl, Pfeiffer, Schauer, Elbert, & Neuner, 2011) and child survivors of disaster (Catani et al., 2009). This approach appears to involve “remaking” of personal narrative. It is conceivable that telling one's life story in a book, art, or film about a traumatic experience may have similar effects on trauma symptoms and/or well-being.

Positive Perspectives on the Self and the Future Survivors of adversity and better-adapting subgroups in studies of resilience frequently report a positive outlook on self, life, and/or the future. These accounts often are retrospective, so it is difficult to know whether overcoming adversity or surviving led to positive feelings and beliefs or whether optimism and hope played a causal role in adaptation or recovery. Longitudinal studies of resilience also suggest that optimism and positive outlooks on life and the future at least accompany resilience, if they do not directly contribute to it (e.g., Werner & Smith, 1992, 2001). Valerie Maholmes published a book in 2014 on fostering resilience in children and families in poverty subtitled: Why Hope Still Matters. In the book, she argues that hope is the “essential mindset” that makes it possible for individuals to show resilience in the face of the adversities associated with poverty. Hope, in Maholmes's view, motivates efforts to overcome adversity and strive for a better future. In the view of Snyder and colleagues (1997), hope is a complex construct combining cognitive and motivational aspects. They created the Children's Hope Scale to measure this construct. In their view, hope is grounded in goals and agency, such that young people with higher hope scores would be expected to have conscious goals, perceive multiple ways to reach to their goals, and be motivated to work and persist in achieving these goals. To date there is very little longitudinal data directly testing the idea that hope or optimism functions directly as a moderator or promotive factor in resilience. However, in a longitudinal study of students age 10 to 18 by Valle, Huebner, and Suildo (2006), hope scores on the Snyder measure showed moderate stability over time and predicted life satisfaction and internalizing symptoms (controlling for starting levels of life satisfaction and symptoms). There also was an interaction of hope with stressful life events. These findings are consistent with the possibility that hope has promotive and protective roles in the context of negative life experiences, although there is a clear need for additional research with more independent assessment of hope, symptoms, life satisfaction, and life events, moving beyond self-reports. Positive ideas and beliefs about self and future are often linked to competence, as noted in many reviews of both competence and resilience (e.g., Kirschman, Johnson, Bender, & Roberts, 2009; Luthar, 2006; Masten & Coatsworth, 1998). More research is needed on the processes that may underlie these associations. Additionally, it is interesting to note that

positive affect, subjective well-being, and feelings of mastery or achievement are associated with resilience, given that several of the major adaptive systems purported to play a role in adaptive function and development have been linked to positive emotions and related neural reward systems. These systems include attachment, mastery motivation, and meaning-making. Reward systems are discussed further in the section on neurobiological processes.

Child-in-Context Resilience Processes In relational developmental systems theory, many aspects of the social context play crucial roles in the interactions that shape development. Resilience science long has emphasized the role of multiple interconnected systems in shaping adaptive capacity. However, during the past decade, research on social systems in resilience has surged, particularly in regard to family and cultural processes. In this section, we highlight advances in research on resilience focused on social contexts, with a particular emphasis on family, school, and cultural systems, where research has been growing.

Family Systems Although resilience has been studied at the level of the family for decades, theory and research on family resilience has not been well integrated with the literature on individual child resilience (Masten & Monn, 2015). This state of affairs is somewhat surprising, given that the study of individual and family resilience share many roots, including general systems theory (von Bertalanffy, 1968). One explanation may be that, until recently, work in family resilience focused more on family therapy and practice; family resilience theory has distinct roots in family systems theory and therapy (Goldenberg & Goldenberg, 2013; Nichols, 2013; Small & Memmo, 2004; Walsh, 2006). In research on individual child resilience, the family always was viewed as a key promoter and protector of development. In the field of family resilience, the focus is shifted to the adaptation of the family as a unit, with challenges that threaten family functions and protective processes that sustain family resilience or facilitate family recovery to full adaptive function (Masten & Monn, 2015; Walsh, 2011). Analogous to the two components for defining resilience in an individual, family resilience can be defined on the basis of family adaptation quality or success and stress or challenges to the family (Patterson, 2002). Proponents of family resilience as a concept highlighted family strengths and protective processes in their theories and practice to help families address or recover from major family stressors, including loss, marital conflict, maltreatment, economic crisis, and divorce (Nichols, 2013; Walsh, 2006, 2011). Families also develop over time; face changing developmental tasks as they form, change, or dissolve; and follow varying pathways of adaptive function (Hawley & DeHaan, 1996; Masten & Monn, 2015). How well a family is doing in the developmental tasks defined at a family level can have profound consequences for children growing up in that family. The resilience of children often depends on the physical and emotional nurturance of their families as well

socialization from the family; if the whole family system is stressed, then the adaptive capacity for a child in that family may well be affected. From a child perspective, families serve as promoters and protectors of development. Families provide emotional security and a sense of belonging, which in turn support learning, exploration, self-efficacy, and social development in many different ways, according to both developmental theory (Bandura, 1997; Bowlby, 1969/1982; Cassidy & Shaver, 2008) and family theory (Goldenberg & Goldenberg, 2013; Walsh, 2006, 2011). Families nurture brain development, language and other cognitive skills, self-regulation skills, and also teach their children about their culture, all of which play key roles in the growing resilience of children throughout the course of childhood (Masten, 2014b). Many of the most dangerous hazards faced by children arise in the family, when a parent is abusive or neglectful, or unfold when the family is unable to protect their children due to death, injury, health problems, separation, or extreme poverty (Cicchetti, 2013; Dubowitz & Poole, 2012; Masten et al., 2015). Family function in crucial family roles mediates many kinds of threats and also moderates (exacerbates or buffers) child experiences in adversity. The Family Stress Model developed by Conger and colleagues (Conger & Elder, 1994; Conger, Reuter, & Conger, 2000) and supported by their empirical work described how economic recessions affect family processes, cascading to influence how children or adolescents fare. Parenting effectiveness can be undermined by depression or by catastrophes affecting the family, such as job loss or a natural disaster that destroys the home or war that leads to forced migration (Masten, 2014b; Masten & Monn, 2015). Recent research also suggests that allostatic load in families can spread to other members of the family, as chronic stress takes a toll both on individuals and on the function of the family as a whole (Repetti, Robles, & Reynolds, 2011). Nonetheless, evidence continues to accumulate that, in times of crisis, an effective parent can mitigate the impact of family illness or economic and other disasters on child function and development (Masten & Monn, 2015; Walsh, 2013). Intervention research that aims to support parents or families in their family roles provides strong evidence that good parenting matters and also that intervention effects can spread to affect other family members (e.g., Fisher, Gunnar, Dozier, Bruce, & Pears, 2006; Patterson et al., 2010; Sandler, Schoenfelder, Wolchik, & MacKinnon, 2011). One of the ways that families work to mitigate stress is by maintaining family routines and regulatory roles, even in times of turbulence or after disaster (Fiese, 2006; Masten & Monn, 2015; Walsh, 2006). Family routines of everyday life, such as family meals or bedtime rituals, can be reassuring or calming to children. In contrast, chaotic family organization with little or no routine is viewed as a marker of family dysfunction that is highly stressful for children (Evans & Wachs, 2009). In multiple respects, families can function as an external regulatory system until children develop their own self-regulatory capacity for handling challenges (Masten & Monn, 2015). Beeghly and Tronick (2011; Tronick & Beeghly, 2011) described how coregulation processes of interaction in infancy shape the development of child self-regulation. There is growing interest in capturing these coregulation processes in research on the family processes

associated with the development of competence and resilience in adversity (Herbers, Cutuli, Supkoff, Narayan, & Masten, 2014). Families also transmit cultural practices and traditions that can promote competence and resilience in their children in multiple ways (Harkness & Super, 2012; Masten & Monn, 2015; Ungar, 2012; Ungar, Ghazinour, & Richter, 2013). These are discussed further in the subsequent section on cultural systems. As the role of families in child resilience grows, there is increasing appreciation of the idea that communities and societies can support family resilience in order to promote or protect the well-being of children at risk of harm. Families can be supported in their role as nurturers of child resilience and moderators of child stress. In the disaster field, for example, there is widespread recognition that family resilience and child resilience are linked, and disaster preparedness and response have changed as a result of this perspective (Masten et al., 2015). Efforts by communities or societies to support child resilience through family-or communitybased efforts also underscore the importance of resilience in larger social systems for both family and child resilience (Becvar, 2013; Conger & Elder, 1994; Masten & Narayan, 2012; Norris, Stevens, Pefefferbaum, Wyche, & Pfefferbaum, 2008; Walsh, 2006).

The Special Roles of Schools in Child Resilience In most societies of the world, education systems play an enormous role in child development. School is the most organized system outside the family where children spend large amounts of their time (Eccles & Roeser, 2012). It is not surprising, therefore, that the school context holds great interest in resilience science. Schools have multiple roles in nurturing resilience (Masten, 2014b). Schools are charged with the development of competence as defined by the community and society supporting the school. Through formal education, children learn and hone many of the psychological skills that will foster resilience later in development, including self-control and problem-solving skills. They acquire knowledge of information and values important for life in their community or culture, along with reading and mathematical skills. Schools can provide positive adult role models, routines, food, health care, opportunities to develop a talent, and other important resources or interactions missing from the lives of some children. Effective schools, classrooms, and teachers have qualities similar to effective parents or families. These include a positive and supportive atmosphere, rules and structure, high expectations, and strong leadership (Bandura, 1997; Cohen, 2013; Doll, 2013; Doll, LeClair, & Kurien, 2009; National Research Council and Institute of Medicine, 2004; Rutter & Maughan, 2002). Schools also provide opportunities for children to experience mastery through guided and scaffolded steps as well as opportunities for relationships with competent adults outside the family. Children spend so much time in school that numerous intervention programs have been designed for implementation in a school context or through after-school programs. One example is the Life Skills Training program developed by Botvin and colleagues (Botvin &

Griffin, 2004; Botvin & Tortu, 1988), a universal school-based intervention designed to prevent substance abuse, with a strong emphasis on building social competence along with specific skills and attitudes for resisting drug use. The Seattle Social Development Project is another school-based intervention with good evidence of efficacy for promoting achievement and reducing risks of antisocial behaviors, based on fostering positive processes of school engagement and relationships (e.g., Hawkins, Kosterman, Catalano, Hill, & Abbott, 2005). Similarly, “Check and Connect,” another efficacious program developed by Christenson and colleagues (Anderson, Christenson, Sinclair, & Lehr, 2004; Christenson et al., 2008) successfully has reduced truancy and school dropout by building school engagement through positive relationships. Recent meta-analyses also suggest that after-school and summer programs are effective, particularly for high-risk children (Durlak et al., 2010; Lauer et al., 2006). There also is considerable research focused on preparing children to be more effective in school through preschool education and prevention programs. Head Start REDI (research based and developmentally informed) was designed to boost the effectiveness of Head Start programs by enhancing the curriculum, teacher training, and materials for families. Results indicate that this program is effective in promoting better social and emotional function as well as academic skills in high-risk preschoolers (Bierman, Domitrovich, et al., 2008; Bierman, Nix Greenberg, Blair, & Domitrovich, 2008). The Chicago School Readiness Project developed by Raver and colleagues (2011) for Head Start classrooms also showed efficacy in a randomized controlled trial, with a specific focus on targeting self-regulation skills as a strategy for improving school readiness. This intervention is a successful example of recent efforts to promote resilience by improving self-regulation skills, particularly during developmental windows of opportunity, including the preschool years (Diamond & Lee, 2011; Zelazo & Carlson, 2012). The preschool years appear to be a developmental window when there is considerable brain plasticity for acquiring and improving self-regulation capabilities that provide a foundation for learning in school. Schools routinely function to nurture human capital and provide resilience-building resources and opportunities for children from typical and disadvantaged backgrounds. However, schools also serve as a powerful symbol of recovery in the aftermath of disasters (American Psychological Association, 2010; Masten & Narayan, 2012; Masten et al., 2015). Reestablishing the routine of school for children is now widely recommended as a crucial restorative step after disasters, symbolizing that recovery is under way. Finally, it should be noted that schools are a major “crossroads” where many important systems of child development intersect and interact, including society, community, culture, family, peers, and individuals. In addition to multiple forms of learning and socialization, many processes of acculturation also occur in the school context. Thus, schools can influence the processes that contribute to peace or conflict among cultures and peoples (Leckman, PanterBrick, & Salah, 2014) as well as the future economic viability of a society (Lundberg & Wuemrli, 2012). Such macro-level processes in turn can alter the future resilience of children and families.

Social Networks and Peers Play, social interaction, and friendships with peers are universal contexts of child development, and social competence with peers is widely recognized as a fundamental developmental task domain beginning in childhood (Hartup, 1996; Hartup & Stevens, 1997; Ladd, 2005; Masten & Coatsworth, 1998). Positive relationships of children with their parents forecast better peer relationships. Once children go to school, their relationships with teachers also are associated with more positive peer relations (Bierman, 2011; Hart, 2012). Interventions at school have successfully targeted peer ecology and interactions, for example, to enhance social-emotional skills (e.g., Durlak, Weissberg, Dymnicki, Taylor, & Schellinger, 2011) or reduce intergroup conflict (Beelmann & Heinemann, 2014). Despite the salience of the peer context in development, there has not been much focus on the potential protective influence of peer relationships in the resilience literature. One exception can be found in the literature on antisocial behavior, where studies show a positive role of peers among high-risk children and youth with respect to delinquent behavior and violence. Having a nondeviant good friend or peer group has shown protective effects (Lösel & Farrington, 2012). Similarly, there is some evidence that high-risk adolescents who report positive peer support have better academic achievement. Gutman, Sameroff, and Eccles (2002) found that peer support moderated risk for math achievement among African-American seventh-grade students, with a significant interaction effect suggesting that peer social and academic support had a protective role at high levels of cumulative risk. Peers also play an important role in the lives of immigrant youth in the classroom. MottiStefanidi (2014) and her colleagues have studied changes in peer acceptance in Greek classrooms with immigrant youth, observing that intergroup contact appears to improve peer acceptance, consistent with bidirectional acculturation effects. Nonetheless, there continues to be a paucity of research on peer roles in resilience, which is surprising, given the large literature on peer relations in developmental science. Peers appear to be studied more often as risk factors or accelerators of risky behaviors than they are studied as protective influences (Dishion & Tipsord, 2011; Steinberg, 2008). Yet, as Dishion and Tipsord (2011) noted, there is no reason to expect that “peer contagion” or similar processes by which peers influence development would only be negative. Although there is evidence of positive roles of peers in the context of risk or adversity, the role of peer systems in resilience, either at a dyadic or larger group level, needs further study.

Cultural Systems Over the past decade, research on resilience grounded in cultural context expanded rapidly, after decades of relative neglect (Masten, 2014a, 2014b; Ungar, 2012; Ungar et al., 2013; Wachs & Rahmann, 2013). There was more research on resilience in diverse cultures as well as efforts to understand the roles of cultural processes in resilience. The Resilience Research Centre in Halifax, Nova Scotia (see resilienceresearch.org), led two major programs to foster cultural perspectives on resilience in multiple countries and cultural contexts. A multinational team developed a new measure of youth resilience and initiated a study of about 1,500 youth

across five continents, using a combination of quantitative and qualitative methods (Ungar, 2008; Ungar, 2012; Ungar & Liebenberg, 2005). This body of work has indicated common and unique promotive/protective processes in different cultures. Common protective influences reported in many cultures include relationships (good parenting, mentoring), cultural belonging (grounded in cultural traditions), and individual characteristics, such as problem-solving or self-efficacy. As part of the Resilience Pathways Project, Theron, Theron, and Malindi (2013) contributed qualitative descriptions of resilience in Basotho youth in South Africa; their work illustrates the importance of Botho in this culture, a philosophy of life akin to Ubunto in other African cultures, where raising children is a collective enterprise and great value is placed on human interdependence, harmony, spirituality, and humanitarian practices. Youth identified as showing resilience by adults in the Basotho community were described as flexible, determined, skilled as communicators, connected to the community, respectful of community values, and accepting of things they could not control, among other qualities. Books and articles focused on risk and resilience in low-and middle-income countries have proliferated over the past decade. These included a volume on early childhood development sponsored by UNICEF in collaboration with the Society for Research in Child Development (Britto, Engle, & Super, 2013) and another focused on risk and resilience in the face of the international economic crisis that was sponsored by the World Bank (Lundberg & Wuermli, 2012). Research on immigrant youth and processes of acculturation also expanded rapidly at the same time. Many books and articles on immigrant youth have focused on resilience, perhaps because this is an issue of global significance for many receiving nations, with many societies depending on the well-being and success of immigrants (e.g., Masten, Liebkind, & Hernandez, 2012; Motti-Stefanidi, 2014). Realizing the Potential of Immigrant Youth (Masten et al., 2012) is an edited volume summarizing recent research in this area. Multicultural societies have a vested interest in preventive and protective interventions to reduce the impact of discrimination on young people and its corrosive collective effect on society itself. There is growing research on strategies for parents, schools, and communities that reduce cultural and intergroup conflicts and foster resilience of young people in the context of discrimination (Evans et al., 2012; Fisher, Busch-Rossnagel, Jopp, & Brown, 2012; Hughes et al., 2006; Masten et al., 2012; Serafica & Vargas, 2006; Ungar et al., 2013; Ungar et al., 2013). Sirin and Gupta (2012) described the resilience of Muslim-American youth in the aftermath of 9/11, a challenging time for multicultural youth of Muslim heritage to navigate, documenting the success of young people who manage to reside happily “on the hyphen.” Motti-Stefanidi and her colleagues, as noted, have studied risk and resilience among immigrant youth in Greece, which has received large numbers of immigrants from the former Soviet Union and neighboring Albania. Using hierarchical linear modeling of longitudinal data, their studies have illustrated changes in self-perceptions and peer interactions over time and the complex ways that classroom context, culture, and individual differences may influence change over time as development and acculturation proceed in concert (Motti-Stefanidi, 2014). Research on resilience in diverse cultures can be a highly challenging endeavor. Panter-Brick,

Eggerman, and colleagues (Eggerman & Panter-Brick, 2010; Panter-Brick, Goodman, Tol, & Eggerman, 2011) have studied resilience in the Afghan people as they suffered through years of political conflict. Despite considerable danger and instability, this team was able to carry out a large study of adolescents and their families, following subgroups over time to gain quantitative data along with rich ethnographic information. Their work emphasized the importance of Afghan values for their resilience, including faith, family unity, service, effort, morals, and honor. Similarly challenging studies of resilience in wars and disasters have been undertaken around the world over the past decade (Masten, 2014a; Masten et al., 2015). These include studies of recovering child soldiers from Sierra Leone (Betancourt et al., 2010, 2013), Palestinian youth engaged in the Intifada (Barber, 2008), and disaster survivors after the 2004 tsunami in the Indian Ocean (e.g., Catani et al., 2009), Hurricane Katrina and the subsequent Gulf oil spill (Osofsky& Ososfky, 2013), the Sichuan earthquake in China (e.g., Luo et al., 2012), and studies under way of the triple disaster in Japan in 2011 (earthquake, tsunami, Fukushima meltdown (Latif, Yeatermeyer, Horne, & Beriwal, in press). Religions are one form of cultural tradition that appears to engage many of the same adaptive systems identified as protective influences in studies of individual resilience (E. Crawford et al., 2006; Masten, 2014b). These systems include relationships, with both spiritual figures and religious leaders and with counselors, rituals to ease the pain of loss and other suffering, belief systems that give people a sense of higher purpose and meaning, and customs of social and material support for members of the religious community and the less fortunate. Religions also codify rules for living and rearing children, passing on values that their members believe will result in good behavior and success in life as defined by the religion. Yet, like other adaptive systems, cultural systems, including religious ideas and practices, can be harmful or dangerous to children and child development, leading to self-destruction or violence, for example. One of the great global challenges facing the world today is addressing the violence to and by children under the banner of radical religious factions or tribal affiliations. Global concerns about the numbers of children affected by war and conflict are giving rise to a growing literature on peace and peacebuilding in childhood (see Leckman et al., 2014).

Other Macro-Level Systems and Resilience Other macrosystems implicated in threatening or supporting child resilience include communities and societal policies as well as mass media and social media (Masten, 2014b). Communities, societies, and media appear to influence resilience in multiple ways, although these processes have received less attention. For example, there is convincing research on the positive effects of economic policies, such as the Earned Income Tax Credit in the United States, showing that tax policies can have measureable benefits for child development (Duncan, 2012). The roles of community preparedness and media have been extensively considered in research and evaluation studies on the impact and recovery of children and families in disasters and war (Masten et al., 2015). Resilience of immigrant youth living in

countries that vary in supports and cultural acceptance of immigrants also has been studied (Fuligni & Telzer, 2012; Masten et al., 2012). Discussion of this expanding and diverse literature is beyond the scope of this chapter. However, it is clear that there is growing interest in research pertinent to resilience that extends across system levels and increasing appreciation of the interconnectedness of resilience among interacting systems. Child resilience depends on family resilience and many other systems that support the development and protect the function of children and families at macro-levels of the socioeconomic and physical ecology. New technologies, including social media, are changing many of these systems in ways that are difficult to study. New horizons of research for resilience science continue to appear.

Neurobiological Promotive and Protective Processes Over the past decade, research on the neurobiology of resilience, long called for and anticipated, finally began in earnest (Cicchetti, 2010, 2013; Charney, 2004; Curtis & Cicchetti, 2003; Kim-Cohen & Gold, 2009; Feder, Nestler, & Charney, 2009; Masten, 2014a, 2014b; Southwick, Litz, Charney, & Friedman, 2011). This surge, which played a central role in propagating the “fourth wave” of resilience science, resulted in large part from major advances in the tools for research at genetic and neurobiological levels of analysis, including such breakthroughs as imaging the brain in action, measuring genes and their functional status, practical measures of stress hormones and other neurobiological variables, and the statistical means to analyze enormous quantities of complex data. Work is proceeding on multiple fronts relevant to resilience. Only a few can be highlighted here. In this section, we describe selected research on the neurobiological processes related to the psychosocial adaptive systems strongly associated with resilience (discussed earlier in this chapter), examples of genetic studies opening new horizons for resilience science, and other promising directions for multiple-levels research on resilience that incorporates biological processes.

Neurobiological Research on Adaptive Systems Implicated in Resilience Major psychological adaptive systems associated with resilience include attachment and caregiving, problem-solving, self-regulation, and mastery motivation. Research at the neurobiological level is beginning to illuminate how these processes develop and protect human function and development. Meaney's groundbreaking work on the biological and behavioral effects of parenting by rat mothers on their pups was already mentioned (Meaney, 2010). Similar research on the role of caregiving in the development of stress, gene expression, and resilience or vulnerability to adversity has been conducted with primates (Stevens, Leckman, Coplan, & Suomi, 2009; Suomi, 2006, 2011). Powerful cross-fostering designs in both animal models (rodents and primates) have shown the effects of parenting and attachment bonds on individual function and

development and gene expression as well as the interplay of individual differences. There is a substantial body of work on social buffering of stress in relation to the hypothalamic-pituitary-adrenocortical (HPA) axis, based on both animal and human studies (Hostinar, Sullivan, & Gunnar, 2014). In addition, research on the “biochemistry of love” by other investigators has shown how oxytocin and vasopressin contribute to processes of social affiliation and attachment behavior in animal models and in humans (Adolphs, 2009; Carter & Porges, 2013, 2014; Porges, 2011). Carter and Porges (2013) have posited that oxytocin plays many roles in adaptation, not only including its well-known roles in birth and lactation, but also in attachment behaviors and neurogenesis. Their ideas have been explicitly tested in prairie voles, animals that show strong attachment bonds. There also have been many advances in understanding how chronic stress affects neurocognitive development, with investigators highlighting the “toxic” effects of severe and prolonged adversity in early development (pre-and postnatal; Gunnar & Herrare, 2013; Shonkoff, 2011). Concomitantly, neurobiological research is emerging on the protective effects of interventions on brain development, theoretically mediated by altering various processes that mitigate negative effects of stress on brain development. Experimental studies of preventive interventions to provide support higher-quality caregiving, nutrition, stimulation, and other key elements of nurturing environments have shown protective effects on brain function and corresponding cognitive skills among high-risk children (e.g., Fisher, Van Ryzin, & Gunnar, 2011; Fox, Almas, Degnan, Nelson, & Zeanah, 2011; Fox, Nelson, & Zeanah, 2014; Nelson et al., 2007; Rutter, Sonuga-Barke, & Castel, 2010). Baker, Salinas, and Eslinger (2012) suggest that school is increasingly viewed as a “neurocognitive-developmental institution.” Research on the neurobiology of self-regulation processes often implicated in resilience also is burgeoning. Examples of hot areas include work on executive function and mindfulness (S. M. Carlson et al., 2013; Zelazo & Carlson, 2012; Zelazo & Lyons, 2011) and stress regulation involving processes linked to the HPA system (Feder et al., 2011; Gunnar & Herrare, 2013; Karatsoreos & McEwen, 2011, 2013; McEwen & Gianaros, 2011). Research on reward systems linked to resilience processes such as mastery motivation also is expanding, although these systems are highly complex. Activation of the dopaminergic reward system, for example, has been linked to positive emotions and motivation associated with resilience as well as to exploratory behavior and learning (Feder, Charney, & Collins, 2011; DeYoung et al., 2011). However, reward systems also play important roles in negative adaptive behaviors, such as maladaptive drug use (Volkow et al., 2011).

Molecular Genetic and Neurobiological Processes Empirical investigations of resilience over the past four decades have examined a wide range of psychosocial correlates and contributors to resilient functioning. In recent years, a number of scientists have urged researchers studying the determinants of resilience to incorporate molecular genetic and neurobiological measures into their investigations of the developmental pathways to resilient functioning (Charney, 2004; Cicchetti, 2010, 2013; Cicchetti and Blender,

2006; Curtis & Cicchetti, 2003; Feder, Nestler, & Charney, 2009; Masten, 2007, 2014; Sapienza & Masten, 2011). Since the onset of the millennium, increasing research has been conducted to discover the contributions that genetics and neurobiology might make to promoting resilience in individuals experiencing severe adversity (Curtis & Cicchetti, 2003; Kim-Cohen & Gold, 2009; Masten & Obradović, 2006). This shift in research from predominantly psychosocial models to those including biological aspects of resilience was spurred by the groundbreaking research on GxE interactions conducted by Avashalom Caspi, Temi Moffitt, and their colleagues (e.g., Caspi et al., 2002, 2003). A high percentage of GxE interaction research conducted to date has demonstrated that child maltreatment is a strong candidate environmental pathogen (Caspi, Hariri, Homes, Uher, & Moffitt, 2010; Karg, Burnmeister, Shedden, & Sen, 2011). Maltreatment is an objectively measured, well-defined stressor, and meta-analyses have shown that robust findings in GxE investigations are more likely to occur when a specific stressor is identified (Karg et al., 2011). Much of the GxE research in the past decade relevant to resilience has focused on how genes may serve as protective factors against developing psychopathology in individuals who have experienced maltreatment. In a longitudinal investigation, Caspi et al. (2002) first documented that the association between child maltreatment and later antisocial behavior was moderated by a functional polymorphism in the promoter region of the gene encoding the monoamine oxidase A (MAOA) enzyme. Specifically, they discovered that maltreated children whose genotype conferred low levels of MAOA expression had significantly higher levels of antisocial behavior in adolescence and adulthood compared to similar-aged maltreated individuals who carried the high-activity version of the MAOA gene. In a subsequent longitudinal study, Caspi et al. (2003) found another GxE interaction involving a functional polymorphism in the promoter region of the serotonin transporter (5-HTT) gene. Individuals who had high stressful life events and/or child maltreatment and who had one or two copies of the 5-HTTLPR “short” allele (S or SS) had more depressive symptomology, higher rates of diagnosable depression, and more suicidality than similarly stressed individuals with two copies of the 5-HTTLPR “long allele” (LL). In both of the Caspi et al. studies (2002, 2003), genetic variation had little or no relation to the mental health outcomes in individuals who had not experienced adversity. Since the publication of Caspi et al.'s (2002, 2003) seminal investigations, there have been numerous GxE studies indicating that genetic variants in a number of genes, including MAOA, 5-HTTLPR, corticotropin-releasing hormone receptor gene (CRHR1), the oxytocin receptor gene (OXTR), and the dopamine receptor D 4 (DRD4) gene, may confer protection against various forms of psychopathology in individuals who have experienced maltreatment or early adversity, such as institutionalization (see, e.g., Bakermans-Kranenburg & van IJzendoorn, 2008; Bradley et al., 2011; Cicchetti, Rogosch, & Oshri, 2011; Dackis, Rogosch, Oshri & Cicchetti, 2012; DeYoung, Cicchetti, & Rogosch, 2011; Gunnar et al., 2012; Hostinar, Cicchetti, & Rogosch, 2014; Kaufman et al., 2004, 2006; Polanczyk et al., 2009). However, it

is important to recognize that the absence of psychopathology is conceptually distinct from competent functioning. Consequently, in order to increase the relevance of GxE research for discovering the determinants of resilience, such investigations must go beyond relying solely on the absence of psychopathology. These studies also should include measures of competence that are likely to be influenced by genotypic variation. GxE studies have focused predominately on how genotypic variants may moderate the effects of environmental pathogens. However, genes also may moderate positive health-promoting environments (Kim-Cohen & Gold, 2009). As such, it is important to conduct GxE research in these environmental contexts. Furthermore, the effects of the interaction of genetic and environmental factors may vary depending on the stage of development (Sroufe, 2013). One of the first published GxE studies that explicitly investigated the relationship between genetic variants and resilience was conducted by Cicchetti and Rogosch (2012). These investigators' research examined whether experiences of child maltreatment in interaction with gene variants of specified candidate genes were related to adaptive functioning. A multicomponent index of adaptive functioning was identified in the sample of maltreated and nonmaltreated school-age children from disadvantaged backgrounds. The genetic variants of HTTLPR, CRHR1, DRD4 -521 C/T,and OXTR were investigated. These candidate genes were specifically selected for their link to behaviors associated with resilience. An adverse main effect on adaptive functioning was consistently observed for maltreatment. Although a main effect was not shown for any of the gene variants investigated, this study identified GxE interaction effects. Nonmaltreated children with the SS genotype variant of 5HTTLPR were significantly more likely to have higher adaptive functioning while maltreated children with the same genetic variant had significantly less adaptive functioning. Analysis of the CRHR1 genotypes presented similar findings in the comparisons between children with zero, one, or two copies of the TAT haplotype. A cross-over interaction presented a differential susceptibility to environmental influences interpretation (Belsky, & Pluess, 2009; Ellis, Boyce, Belsky, Bakermans-Kranenburg, & van IJzendoorn, 2011) for those children with zero copies of the TAT haplotype in that they exhibited better adaptation in the context of nonmaltreatment and worse adaptation with maltreatment. Analysis of the OXTR gene variants contrasting the GG with the AA or AG genotypes produced similar results, but they were only marginally significant. Analyses of the DRD4 -521C/T single nucleotide polymorphism (SNP) revealed a different pattern. The interaction (contrasting the TT genotype group with CC or CT genotypes) suggested that for this candidate gene, variations in genotype were more important for nonmaltreated children. Nonmaltreated children with the TT genotype were more likely to exhibit the highest level of adaptive function. In the context of maltreatment, which had a main effect on adaptive function, genotype variation did not appear to matter. In further analyses, Cicchetti and Rogosch (2012) examined the relation between number of maltreatment subtypes as well as specific subtypes of maltreatment in relation to DRD4 -512C/T genotype in predicting adaptive function. Maltreated children who had experienced three or four subtypes fared dramatically worse with the TT genotype, whereas nonmaltreated

children with this genotype fared the best. These results are congruent with a differential susceptibility role of the TT genotype of DRD4 -521C/T in relation to environmental variations in maltreatment. Cicchetti and Rogosch (2012) identified different gene variants as differentially important through contrasting maltreated and nonmaltreated children in relation to adaptive function. Genotype was not generally related to adaptive function unless context was taken into account. Maltreatment, in contrast, appeared to have powerful negative main effects for adaptive function. It is important to note that despite the multidomain assessment of adaptive functioning and the large, well-defined sample of maltreated and nonmaltreated children, a limitation of this study was its cross-sectional nature. Given that resilience is a dynamic developmental construct that is not immutable, future investigations of gene–environment interaction and resilience should have a prospective, longitudinal, and developmentally sensitive design (Luthar, 2006; Luthar et al., 2000). It also is important to acknowledge that other biological systems are negatively impacted by abuse and neglect (Cicchetti & Toth, 2015; DeBellis, 2001, 2005). Differences in brain structure and function (Hart & Rubia, 2012; McCrory et al., 2010), stress neurobiology, (Cicchetti, Rogosch, Gunnar, & Toth, 2010; Doom & Gunnar, 2013; Heim et al., 2008; Trickett et al., 2008) and immunological processes (Danese et al., 2008; Miller et al., 2011) are not innate features of children but rather are consequences of child maltreatment. Maltreatment experiences also may influence other developmental processes, including epigenetic processes involving the extent to which genes are expressed (Tyrka, Price, Marsit, Walters, & Carpenter, 2012); however, these effects are not examined when focusing exclusively on GxE interaction. Future investigations of molecular genetic contributions to resilience also should pay more attention to the process of development (D). These GxE, GxD, and DxExD studies could discover how a particular genotype may operate differently depending on developmental period or environmental period or environmental context (see Sroufe, 2007). Moreover, genotypic variation for genes other than those utilized in the Cicchetti and Rogosch (2012) investigation may be related to higher resilient functioning in maltreated children. Thus, broader approaches to genetic influences associated with resilience in maltreated children must occur.

Epigenetic Studies of Resilience Epigenetic research pertinent to resilience is rapidly expanding, although it likely will take some time for this knowledge base to mature. The field of epigenetics has demonstrated that DNA is not the static entity it was once thought to be (Mill, 2011). Epigenetics is the study of heritable changes in gene function that do not alter the DNA sequence but instead provide an extra layer of transcriptional control that regulates gene activity and plays an important role in the acute regulation of genes in response to environmental changes. Research on epigenetic processes hold high promise both for basic research investigations and translational intervention research on the determinants of resilience, providing a potential mechanism by

which environmental experiences (natural or in the form of intervention) may exert their impact on gene expressions that can last throughout the life span or even across generations (Lenroot & Giedd, 2008). Epigenetic mechanisms may be a realistic target for intervention due to their reversibility (Ramchandahl, Bhattacharya, Cervoni, & Szyf, 1999). However, some epigenetic changes are long lasting across the course of development. Moreover, some of these methylation changes can be transmitted across generations (Roth, 2013). Research on early adversity, such as child abuse and neglect and living in impoverished environments, has demonstrated that there are individual differences in DNA methylation patterns. Methylation assays can be conducted genome-wide or at site-specific regions of candidate genes whose functional properties are known. For example, several investigations have used a candidate gene approach to examine the effects of child maltreatment on methylation. Epigenetic regulation is a viable candidate mechanism through which caregiving behavior, including child maltreatment, may exert longterm effects on HPA axis activity and neuronal function (Szyf & Bick, 2013). Candidate genes utilized in this regard include the serotonin transporter gene (5-HTT), the glucocorticoid receptor gene (NR3C1), and the binding protein 5 gene (FKBP5). These studies have found that child maltreatment was associated with increased methylation in both the serotonin transporter and the glucocorticoid receptor genes (see Yang et al., 2013). Increased methylation (hypermethylation) is commonly associated with decreases in gene expression. Decreased methylation (hypomethylation) is associated with increases in gene expression— often of genes that should not be expressing. Epigenetic mechanisms may help to elucidate pathways of risk and resilience and illuminate how interventions work. The dynamic nature of epigenetic processes demonstrates that genetic influences on behavior are responsive to experience and the environmental surround. Protective factors in the environment may bring about alterations in gene expression and thereby set in motion a cascade of positive adaptation to stress that is promotive of resilience (Bowes & Jaffee, 2013; Masten & Cicchetti, 2010c).

Multiple-Levels in Developmental Reorganization and Plasticity As emphasized throughout this chapter, several cogent and extensive summaries of resilience theory and research have been published that have explicated the need to examine the processes leading to resilience from multiple levels of analysis, in particular from the level of brain and neurobiological functioning (Cicchetti, 2010, 2013; Cicchetti & Blender, 2006; Feder, Nestler, & Charney, 2009; Luthar, Cicchetti, & Becker, 2000; Masten, 2001, 2007, 2014b). Advocates of a self-organizing systems theory viewpoint of neurobiological and psychological development contend that individuals actively participate in the creation of meaning by structuring and restructuring experience through self-regulated mental activity (Cicchetti & Tucker, 1994; Mascolo, Pollack, and Fischer, 1997). Early experience and prior levels of adaptation neither doom the individual to continued maladaptive functioning nor inoculate the individual from future problems in functioning. Change takes place in a system as new needs and environmental challenges destabilize the existing organizations, necessitating the emergence of new organizations that may prove to be more adaptive than the preexisting

ones in the current context. The reorganizations that occur both within and between developmental domains (e.g., between emotion and cognition) provide critical opportunities for changes in capacity for adaption, thus influencing prospects for resilience (Cicchetti & Tucker, 1994). The reorganization that takes place is often dramatic, with observable changes in the neural substrate that are translated into changes observable at the behavioral level (Cicchetti, 2002; Curtis & Nelson, 2003). The brain can change in a number of ways, including: (a) endogenous changes that take place in the organization of the remaining intact circuits in the brain; (b) the generation of new neural circuitry; (c) dendritic outgrowth and the formation of synapses; (d) the strengthening and weakening of some neural circuits through the processes of synaptogenesis and synaptic removal, respectively; (e) the sprouting of new axons, regenerating old ones, or elaborating their dendritic surfaces; and (f) the generation of new neurons and glia to replace at least some lost neurons (Cicchetti, 2015; Kolb & Gibb, 2001). Such central nervous system changes that are the hallmarks of plasticity can occur on one or more levels of analysis, including molecular, cellular, neurochemical, and anatomical and at the level of brain systems (van Praag, Kempermann, & Gage, 2000; Nelson, 1999). Plasticity also plays a prominent role in neurodevelopmental processes (Cicchetti & Cannon, 1999; Cicchetti & Walker, 2003; Cowell, Cicchetti, Rogosch, & Toth, 2015). Biological factors influence psychological processes; however, social and psychological experiences also exert actions on the brain by feeding back upon it to modify gene expression and brain structure, function, and organization (Cicchetti & Tucker, 1994; Eisenberg, 1995; Kandel, 1998; Kolb & Wishaw, 1998). It has also been discovered that alterations in gene expression induced by learning and through social and psychological experiences produce changes in patterns of neuronal and synaptic connections and thereby in the functioning of nerve cells (Kandel, 1998). Such neuronal and synaptic modifications not only exert a prominent role in initiating and maintaining the behavioral changes that are provoked by experience but also contribute to the biological bases of individuality, as well as to individuals being differentially affected by similar experiences, regardless of their positive or negative/adverse valence (Kandel, 1998). Because the brain is only diffusely structured by gene-driven processes and the eventual differentiation of the brain is very reactive to the individual's active coping in a particular environment, the normal, abnormal, and resilient outcomes of this developmental process should encompass a diverse range of neural, synaptic, and associative networks that are the physiological determinants of many possible individual psychological organizations (Cicchetti, 2007; Kandel, 1998). Diversity in process and outcome are hallmarks of the developmental psychopathology perspective (Cicchetti, 1990; Cicchetti & Rogosch, 1996; Sroufe, 1989). The application of equifinality to resilience requires that researchers must understand that the maintenance or improvement in a system's function, just as is true with its breakdown, can occur in many ways, even in the face of significant adversity. In particular, because the organizational model of development that undergirds much research in developmental psychopathology and resilience takes into account environment-organism interactions and transactions, it becomes quickly apparent that a variety of developmental progressions may eventuate in maladaptive or

adaptive outcomes (Luthar et al., 2000). Clearly, biological and psychosocial factors both can play a role in the pathways to diverse outcomes. Furthermore, the relative contributions of biological and psychosocial contributions to these diverging pathways will vary among individuals. Likewise, the concept of multifinality alerts resilience researchers to the fact that individuals may begin on the same major developmental trajectory and, as a function of their subsequent “choices” or experiences, exhibit very different patterns of maladaptation or adaptation (Cicchetti & Rogosch, 1996; Sroufe, 1989; Sroufe, Egeland, & Kruetzer, 1990). Pathways of psychopathology or resilience are influenced by a complex matrix of the individual's level of biological and psychological organization, current experience, social context, timing of the adverse event(s) and experiences, and developmental history. As the developmental process proceeds, neural plasticity appears to be guided by events that provide information to be encoded in the nervous system (Greenough, Black, & Wallace, 1987). These experience-dependent processes involve the adaptation of the brain to information that is unique to the individual. An important central mechanism for experiencedependent development is the formation of new neural connections. Because all individuals encounter distinctive environments, each brain is modified in a singular fashion. This variability in experience-dependent plasticity is likely to emanate from existing differences between neural subsystems in rates of maturation, the extent and timing of redundant connectivity, and the presence of the neuromodulators and receptors that are critical for neural modifiability. Unlike the case with experience-expectant processes, experience-dependent processes do not take place within a stringent temporal interval because the timing or nature of experience that the individual engages or chooses cannot be entirely and dependably envisioned (Black, et al., 1998; Cicchetti, 2002). Given that experience-dependent processes can occur throughout the life span, social interactions, behavioral interventions, and pharmacotherapy, in conjunction with individuals' self-righting tendencies, have the capacity to repair brains that are afflicted with disorders (Black et al., 1998; Cicchetti, 2002). There is growing evidence that successful intervention modifies not only maladaptive behavior but also the cellular and physiological correlates of behavior (Cicchetti, Rogosch, Toth, & Sturge-Apple, 2011; Dozier et al., 2008; Fisher, Stoolmiller, Gunnar, & Burraston, 2007; Kandel, 1998). Such interventions are critical to implement because abnormal perturbations at a particular stage of brain development may lead to the development of aberrant neural circuitry and often compound themselves into relatively enduring forms of psychopathology (Arnold, 1999; Nowakowski & Hayes, 1999). In self-organizing brain development, some regions of the brain serve to stabilize and organize information for other areas, whereas other regions utilize experience to fine-tune their anatomy for optimal function (Singer, 1995). In this manner, individuals can benefit from the interaction of genetic constraints and environmental information to self-organize their highly complex neural systems. Thus, children manifesting resilience, who develop in a positive and adaptive ways even though they have experienced adversity, may be likely to function adaptively in a wide array of environments. Children play an active role in seeking and receiving the

experiences that are developmentally appropriate for them. Through experience-dependent processes, children who function adaptively are likely to seek out experiences that serve to modify and/or protect their brain development in ways that foster future adaptive developmental outcomes and generate resilience. In effect, the individual is helping to create his or her capacity for adaptive development. At one level, different functional units of the brain may attempt to compensate for dysfunctions in another functional aspect, and at another level, the organism may seek out new experience in areas where it has strength (Black et al., 1998). Because experience-dependent plasticity is a central feature of the mammalian brain, neither early brain anomalies nor aberrant experiences should be viewed as determining the ultimate fate of the organism (Cicchetti & Tucker, 1994).

Normative Biological Processes in Resilience The biological processes implicated in resilience and discussed briefly in this chapter are in fact normative processes. Neural plasticity, for example, is an inherent property of the central nervous system. This observation represents an interesting parallel to Masten's (2001) suggestion that resilience is common and typically emerges through the operation of ordinary human adaptational systems. Likewise, at the biological level of analysis, normative processes may mediate most pathways of resilience, as long as these systems are functioning within normal parameters. This viewpoint underscores the importance of the interaction of normative systems at all levels of analysis in the promotion of resilience.

Empirical Studies on the Biological Contributors to Resilience In 2004, Charney proposed an integrative model of vulnerability and resilience that encompassed the neurochemical response patterns to acute stress and the neural mechanisms mediating reward, fear conditioning and extinction, and social behavior. Charney identified 11 potential neurochemical, neuropeptide, and hormonal mediators of the psychological response to extreme stress. He concluded that the neural mechanisms of reward and motivation (e.g., hedonic capacity, optimisim, and learned helpfulness), fear responsiveness (e.g., engaging in effective behaviors despite fear), and altruism were relevant to the personality/character traits associated with resilient functioning. It will be important to conduct empirical research relating the neurobiology of these systems to a multidomain assessment of resilient functioning (Masten, 2014b). In the context of normative conditions, the connections of the medial prefrontal cortex with the amygdala are immature during childhood and become more mature during adolescence. Gee et al. (2013) found that children who were previously institutionalized, thereby experiencing early maternal deprivation, exhibited an early emergence of mature medial prefrontal cortex connectivity and resembled the adolescent/adultlike phenotype. This pattern of connectivity, which was mediated by the cortisol hormone, may confer some degree of enhanced emotion regulation because maternally deprived children with adolescent/adultlike phenotypes manifest less anxiety than their counterparts with immature phenotypes. Gee et al. conjectured that the accelerated amygdala-prefrontal development may serve as an ontogenetic adaptation in individuals in response to early adversity.

Peterson et al. (2014) used fMRI to compare brain function during the performance of an attentional, self-regulatory task. They studied a large community sample of multigenerational families enrolled specifically to be at either high or low risk for developing major depressive disorder (MDD). The sample was comprised of individuals who were at higher or lower risk for developing MDD based on being a biological offspring of persons with MDD or a control sample of persons drawn from the community who had no discernible lifetime illness. Peterson and colleagues discovered a risk endophenotype that included greater activation of cortical attention circuits and a protective endophenotype that included a greater activation of the dorsal anterior cingulate cortex. Thus, neural systems were identified that increased the risk for depression, protected individuals from the illness, and that endured following illness onset. Few multilevel, multidomain empirical studies of the predictors of resilient functioning have been conducted. Curtis and Cicchetti (2007) investigated whether the positive emotionality and increased emotion regulatory ability associated with resilient functioning would be related to greater left frontal electroencephalogram (EEG) activity. Hemispheric asymmetries in EEG activation index a neural system that has emotion-specific influences such that the two hemispheres of the cerebral cortex are differentially involved in emotions (Davidson, 2000). The left hemisphere has been shown to be associated with positive emotions/approach behavior, and the right hemisphere has been linked to negative emotions/withdrawal behavior. Positive emotion and good emotion regulatory abilities have consistently been associated with resilience (Bonanno et al., 2007; Masten, 1986). Thus, the potential connection of hemispheric EEG asymmetry with resilience lies in its common linkage with emotion and emotion regulation. The investigation of a neural-level phenomenon such as hemispheric EEG asymmetry in the context of resilience underscores the point that no one single characteristic will be ascendant in resilience processes of adaptation over the course of development. EEG asymmetry across central cortical regions distinguished between children with high and low resilience manifested in adaptive functioning, such that left-hemisphere activity characterized those maltreated children who were adapting well based on a multilevel, multidomain competence composite index (e.g., effective peer relations, adapting successfully to school, low levels of depressive symptomatology, low externalizing and internalizing psychopathology). In addition, a behavioral measure of emotion regulation, based on 35 hours of observation of the children, also independently contributed to the prediction of resilience in maltreated and nonmaltreated children. In a multilevel investigation of another sample of low-income maltreated and nonmaltreated children, resilient functioning in relation to the regulation of two stress-responsive adrenal steroid hormones, cortisol and dehydroepiandosterone (DHEA), as well as the personality constructs of ego resiliency and ego control, was examined. A composite measure of adaptive success was utilized in this study of resilience that included multidomain, multimethod, multiinformant assessments of competent peer relations, school success, and low levels of internalizing and externalizing symptomatology. Cicchetti and Rogosch (2007) found that the personality characteristics of ego resiliency and moderate ego overcontrol, and the adrenal steroid hormones associated with stress (i.e., cortisol and DHEA) made independent and noninteractive contributions to resilience. Although operating at different levels of analysis,

behavioral/psychological and biological factors each made unique contributions to resilience. Cicchetti and Rogosch (2007) also discovered that higher morning levels of cortisol were related to lower levels of resilience evident in the functioning for the nonmaltreated children. High basal cortisol may indicate that nonmaltreated children are experiencing greater stress exposure and consequently are constrained in their ability to adapt competently. Physically abused children with high morning cortisol had significantly better functioning than physically abused children with lower levels of morning cortisol. The positive role of increased cortisol for physically abused children is divergent from the more general pattern of higher cortisol being related to lower competence in functioning that was discovered in the nonmaltreated and sexually abused children in this investigation. Prior research on neuroendrocrine regulation has indicated that physically abused children generally exhibit lower levels of morning cortisol secretion (Cicchetti & Rogosch, 2001a), especially those children who experience physical abuse early in life (Cicchetti, Rogosch, Gunnar, & Toth, 2010; Cicchetti, Rogosch, & Oshri, 2011). It may be that the subgroup of physically abused children who had elevated cortisol to cope with stressful vicissitudes in their lives were demonstrating a greater capacity for positive adaptation that signifies resilience in progress. In contrast, the larger subgroup of physically abused children with lower levels of morning cortisol may have developed hypocortisolism over time in response to chronic stress exposure. As a result, these children may have a diminished capacity to mobilize the HPA axis to promote positive adaptation under conditions of ongoing stress. Finally, maltreated children with good adaptive functioning exhibited a unique atypical pattern of a relative DHEA diurnal increase. Maltreated children who have evinced elevated DHEA over the course of the day may be better equipped to cope with the demands of high chronic exposure to stress and to adapt competently. In contrast, the nonmaltreated children who functioned well did not exhibit the pattern of diurnal DHEA increase; instead, they displayed the lowest levels of DHEA across the day. DHEA has been shown to generate effects that are associated with resilience. For example, DHEA counteracts the harmful effects of hypercortisolism, thereby potentially conferring neuroprotection to the brain (Charney, 2004). The experience of childhood maltreatment has been shown to be related to changes in the structure of the brain (Hart & Rubia, 2012; McCrory et al., 2010; Teicher, Tomoda, & Andersen, 2006). Little is known about the neurobiological contributors that protect against the negative sequelae that commonly occur with child maltreatment (Cicchett & Toth, 2015). Van der Werff et al. (2013) obtained resting-state fMRI scans in a sample of 33 adults. Twenty-two of these individuals had been maltreated, and the remaining 11 served as controls. Eleven of the previously maltreated adults formed the resilient group, whereas the other maltreated individuals form the vulnerable group. Van der Werff and colleagues sought to identify restingstate functional connectivity patterns specific for resilience to the effects of child maltreatment. Resting-state functional connectivity patterns were discovered that were specific for individuals in the resilient groups. The regions of interest that were implicated were related functionally to declarative memory and the processing of emotional stimuli. Sheinkopf et al. (2007) examined the predictors of better outcomes in children with prenatal

cocaine exposure (CE). A 15-item risk index of prenatal and postnatal experiences was developed for children with and without CE. Preschool cognitive outcomes, problem behaviors, and adaptive behaviors were assessed. Vagal tone (VT) was assessed on two occasions—at an infant exam at 1 month and toy exploration at 36 months. Children were classified as having consistently high, consistently low, or fluctuating VT. Moreover, children were also classified as high versus low risk. High-risk children were found to have lower IQ scores, more problem behaviors, and lower adaptive behavior ratings than low-risk children. The investigators also conducted a risk-by-VT stability interaction that indicated that the highrisk children who had stable low VT had higher ratings of adaptive behaviors at 36 months. These results were viewed as consistent with theory that links reduced VT during tasks to adaptive regulation and suggest that such regulatory function may serve as a protective factor for children with prenatal CE. Nigg and colleagues (2007) classified children as showing resilience if they avoided developing attention-deficit/hyperactivity disorder (ADHD), oppositional defiant disorder (ODD), or conduct disorder (CD) in the face of family adversity. Neuropsychological response inhibition and a composite catecholamine genotype risk score were chosen to be potential protective factors because of their potential relevance to prefrontal brain development. Resilient children showed more effective response inhibition. Moreover, the high-risk genotype index that was comprised of allelic variants of three genes chosen that express in the prefrontal cortex (i.e., dopamine transporter, dopamine D4 receptor, and noradrenergic alpha-2 receptor) was a reliable indicator of resilience with respect to the development of ADHD and CD but not ODD in the face of significant psychosocial adversity. Thus, neurobiological and genetic factors related to prefrontal cortex development may serve as protection by enabling children to avoid developing ADHD and CD in the context of significant adversity.

Avoiding Biological Reductionism Future research on the determinants of resilience should increasingly contain both biological and psychological measures. We want to stress, however, that our viewpoint does not reduce resilience to biology, let alone to a unitary biological variable. A multiple levels of analysis perspective on resilience should not be misinterpreted as equating resilience with biology. The incorporation of a biological perspective into research on resilience still requires adherence to a dynamic, transactional view that acknowledges and respects the importance of context and the interplay through interactions across multiple levels of function. Biological and psychological domains are both essential to include in basic research on resilience and in resilience-promoting interventions. If we are to grasp the true complexity of the concept of resilience, then we must investigate it with a commensurate level of complexity.

Translational Research and Implications Resilience science has had transformative effects on interventions and practice aiming to promote better outcomes or function among children at risk for many different reasons. The initial stage of transformation occurred at a conceptual level as many applied fields shifted

toward strength-based models and away from deficit-oriented disease models and then rapidly infused more positive models, measures, and goals into their practices (Masten, 2011). This shift can be observed in: school psychology and school counseling (Akos & Galassi, 2008; Masten et al., 2008); family therapy (Goldenberg & Goldenberg, 2013; Walsh, 2006, 2011); disaster preparedness (Masten et al., 2015); training and services for military families (Cozza & Lerner, 2013; Masten, 2013; Park, 2011); positive youth development and out-of-school programming (Lerner, 2009); child welfare reform (Flynn et al., 2006, Masten, 2006); social work (Anthony, Alter & Jenson, 2009); medicine (Szanton & Gill, 2010; Southwick et al., 2011); and the practices of humanitarian agencies (Ager, 2013; Lundberg & Wuermli, 2012; Masten, 2014a). Luthar and Cicchetti (2000) observed that research on resilience has substantial potential to guide the development and implementation of interventions for facilitating the promotion of adaptive function and development in diverse high-risk populations that have experienced significant adversity. Prevention scientists have contended that it makes far more sense to promote the development of adaptive functioning as early as possible in the course of development than it does to implement treatment strategies designed to repair existing disorders among high-risk individuals. From its inception as an integrative science, developmental psychopathology spurred investigators to discover the most effective means of preventing and ameliorating maladaptive and pathological outcomes (Cicchetti, 1984; Rutter & Garmezy, 1983; Sroufe & Rutter, 1984). An early stated goal of developmental psychopathology was to help reduce the schism between empirical research and its application to the treatment of individuals struggling with adversity and mental disorders (Cicchetti, 1990). In reflecting on the past achievements and future promises of developmental psychopathology, Cicchetti and Toth (2009) stated that “one of developmental psychopathology's potential contributions lies in the heuristic power it holds for translating facts into knowledge, understanding, and practical application” (p. 17). Knowledge of developmental norms, appreciation of how developmental level may vary within same age groups, sensitivity to changing meaning that problems have at different developmental levels, attention to the effects of developmental transitions and reorganizations, and understanding of the factors that are essential features to incorporate into the design and implementation of preventive interventions all may serve to enhance the potential for optimal intervention efficacy and the promotion of resilience. The principles of developmental psychopathology (Cicchetti, 1993; Cicchetti & Toth, 2009) lend themselves to fostering translational research that has implications for society and for policymakers. The very subject matter of developmental psychopathology necessitates thinking clearly about the implications of the work and devising intervention strategies that will promote competent functioning and positive developmental pathways in challenging contexts (Masten, 2014b). In this vein, efforts to prevent the emergence of psychopathology or to ameliorate its effects also can be informative for understanding processes involved in the development of resilience and adaptive functioning (Cicchetti & Blender, 2006; Masten, 2014b). For example, if the developmental course is altered as a result of a preventive intervention and the risk for negative outcomes is reduced, then prevention research helps to specify processes that are

involved in the development of psychopathology or other outcomes. As such, prevention research can be conceptualized as true experiments in altering the course of development, thereby providing insight into the etiology and pathogenesis of disordered outcomes and to the promotion of resilience (Cicchetti & Hinshaw, 2002; Howe, Reiss, & Yuh, 2002). The experimental nature of randomized clinical trials (RCTs) provides an unparalleled opportunity to make causal inferences in resilience research. Independent variables manipulated in RCTs may be several steps removed from underlying etiologic factors because such trials are primarily concerned with alleviating suffering and promoting competence. Nonetheless, careful research design and assiduous measurement of ancillary process variables through which intervention effects may occur can shed light on the theory-driven mechanisms underlying pathways of pathological and positive adaptation (Cicchetti & Hinshaw, 2002).

Intervention Research Based on Resilience Models Reviewing even the past decade of intervention research based on resilience models is beyond the scope of this chapter. However, selected examples can highlight major trends in resiliencefocused intervention science. Throughout this chapter we have noted examples of interventions that target promotive or protective processes in order to foster positive development or recovery in children exposed to adversity. Examples of resilience-based interventions include carefully controlled experiments to test programs designed to: promote parenting skills among parents with children or youth at risk for antisocial and other high-risk behavior (e.g., Brody et al., 2004; Patterson et al., 2010); improve parenting in families at risk for child maltreatment (e.g., Toth & Gravener, 2012) or in foster parents (e.g., Dozier et al., 2008; Fisher et al., 2006); implement foster care or adoption as an alternative to institutional care (Nelson et al., 2007; Rutter, Sonuga-Barke, & Castle, 2010): and improve executive function skills among disadvantaged preschoolers (e.g., Raver et al., 2011). The best of these studies provide a theory of change that delineates the process targeted for change. Executive functions were the mediator targeted by Raver et al. (2011) in their intervention for high-risk preschool children. Their intervention improved school readiness, and their study further indicated that this change was mediated, as hypothesized, by improvements in cognitive self-control skills. Similarly, experiments to promote resilience in families with children at risk for maltreatment from the parent have targeted the quality of parent-child interaction and the attachment relationship, showing better outcomes in children. Illustrative randomized control trials are highlighted in the next section. Interventions that showed cascade effects were particularly noteworthy over the past decade. In their intervention to promote better outcomes in foster children, discussed further later, Fisher et al. (2006) intervened at the parenting level and measured the outcome in the biological stress system of the child, finding that the intervention had a normalizing effect on the child's HPA system. The Strong African American Families Program (Brody et al., 2004; 2006) intervened with families to alter the risky behaviors of their adolescents. The Oregon

Social Learning Center group, led by Patterson, Forgatch, and DeGarmo (2010), showed spreading effects of their Oregon model of parent management training (PMTO) on multiple members of the family over multiple domains over time. Intervention studies also are beginning to test genetic moderating effects of response to intervention. In their intervention with African-American families, Brody et al. (2009, 2013) tested gene by intervention effects of intervention in their randomized controlled trial. They found evidence of genetic moderating effects, some of which are congruent with a differential susceptibility hypothesis (Brody et al., 2013). Such research on genetically (or epigenetically) moderated treatment effects in resilience science are in the early stages but are likely to be an important area of growth in the coming decade.

Randomized Controlled Trials Targeting Parent-Child Relationships Interventions successfully targeting the quality of care-giver-child relationships in RCT designs with high-risk families offer powerful evidence of promotive or protective roles of this relationship in child resilience. Groundbreaking RCT experiments have been conducted with children and their caregivers at risk due to maternal depression, maltreatment, and foster care. In a study of families at risk due to maternal depression, Cicchetti, Toth, and Rogosch (1999) conducted an RCT with families of 20-month-old toddlers. The mothers suffered from Major Depressive Disorder (MDD), which poses serious risks to child development (Cicchetti & Toth, 1995). Half of these youngsters and their mothers with MDD were randomly assigned either to toddler-parent psychotherapy (TPP; Lieberman, 1992) or to an enhanced community standard control group. A third group was comprised of demographically comparable toddlers whose mothers had no past or present MDD or other psychopathology (normative controls; NC). The intervention lasted 16 months. At baseline, there were no differences found in attachment security between the two groups with mothers who had MDD; both groups of MDD offspring had significantly more insecurity than the toddler offspring of well mothers. At the conclusion, Cicchetti et al. (1999) found that the TPP intervention increased attachment security in the MDD group receiving this intervention to a percentage comparable to that seen in the toddler offspring of well mothers. The MDD group with the community standard of care did not show improvement in attachment security. The results of this investigation are noteworthy; they were one of the first to suggest that attachment security is malleable and can be modified through the provision of an intervention informed by attachment-theory. In a later and larger RCT investigation of mothers with MDD and their toddlers by Toth, Rogosch, Manly, and Cicchetti (2006), attachment security was assessed in the Strange Situation (Ainsworth, Blehar, Waters, & Wall, 1978) rather than with the Attachment Q-Sort (ASQ) that had been used in the 1999 study by Cicchetti and colleagues. The Strange Situation is the gold standard measure of attachment security in the field. Moreover, security of attachment as assessed in the Strange Situation has been shown to relate to a number of later positive outcomes (e.g., autonomous self-development, effective peer relations, ego-resiliency,

fewer behavior problems, etc.; Cicchetti & Roisman, 2011; Sroufe, Egeland, Carlson, & Collins, 2005). To further elucidate the impact of TPP on attachment organization, Toth et al. (2006) conducted an RCT with mothers who had MDD and their toddlers. The toddler offspring of MDD mothers were randomly assigned to either TPP or to a community standard intervention control. A third group of mothers who had never experienced psychopathology and did not differ from the two groups with MDD (intervention and control) demographically served as a normative comparison group (NC). The TPP intervention lasted approximately 18 months. At baseline, toddlers in the two groups with mothers who had MDD were significantly more likely to be insecurely attached to their mother than were toddlers in the NC group. The percentages of Type D disorganized attachment at baseline were 45%, 37%, and 20% for the MDD control, the MDD TPP intervention group, and the NC groups, respectively. At postintervention, statistically significant differences were obtained in security of attachment between the mothers with MDD who received TPP intervention (67%) compared with the community standard control (17%) group. Furthermore, the TPP intervention group had significantly more attachment security than the NC group (48%). The percentages of Type D disorganized/disoriented attachment at postintervention were 11%, 41%, and 21% on the TPP, community standard, and NC groups, respectively. The results of the TPP intervention for toddlers of mothers with MDD provide compelling support for the potential malleability of attachment insecurity. The TPP intervention was effective not only at modifying attachment insecurity but also in maintain existing security in the toddler offspring of depressed mothers. Conversely, offspring of mothers with MDD not receiving the TPP intervention were less likely than intervention offspring to maintain secure attachments at the postintervention assessment. Thus, even if offspring of mothers with MDD are securely attached, the parenting milieu and the emotional climate of the home that accompany maternal depression, even once an active episode has remitted, may mitigate against the continuance of their secure attachment. The malleability of insecure and disorganized attachment among infants from maltreating families also was investigated through an RCT (Cicchetti, Rogosch, & Toth, 2006). Findings from research on the effects of maltreatment on infant attachment were incorporated into the design and evaluation of the intervention (V. Carlson, Cicchetti, Barnett, & Braunwald, 1989; Cicchetti & Barnett, 1991; Cyr, Euser, Bakermans-Kranenburg, & van IJzendoorn, 2010). Oneyear-old infants from maltreating families and their mothers were randomly assigned to one of three intervention conditions: (a) child-parent psychotherapy (CPP), (b) psychoeducational parenting intervention (PPI), and (c) community standard (CS) controls. A fourth group of infants from nonmaltreating families and their mothers served as an additional low-income normative comparison (NC) group. At baseline, mothers in the maltreatment group, relative to the nonmaltreatment group mothers, reported greater abuse and neglect in their own childhoods, more insecure relationships with their own mothers, more maladaptive parenting attitudes, more parenting stress, and lower family support, and they were observed to evince lower maternal sensitivity. Infants in the maltreatment groups had significantly higher rates of disorganized attachment than infants in the NC group. At postintervention follow-up at age 26

months, children in the CPP and PPI groups demonstrated substantial increases in secure attachment, whereas increases in secure attachment were not found for the CS and NC groups. Moreover, disorganized attachment continued to predominate in the CS group. Following the tenets of an organizational perspective, it is expected that these maltreated youngsters, now that they are traversing a more positive developmental trajectory, will be more likely to continue on an adaptive pathway and successfully resolve future salient developmental tasks (Sroufe, 1983). The preventive interventions have demonstrated that behavioral plasticity is possible, at least in the early years of life. The parenting education intervention and the CPP intervention were both found to be equally efficacious in fostering secure attachment and in reducing disorganized attachment in infants from maltreating families. However, when examining the sustained efficacy of these intervention models, only CPP was found to be efficacious in promoting quality of attachment over time (Pickreign Stronach, Toth, Rogosch, & Cicchetti, 2013). These results suggest that, in cases of extremely maladaptive parenting such as child maltreatment, more intensive models of intervention that go beyond parent skills training may be necessary as increasing developmental complexity unfolds over time. The extremely high stability of attachment insecurity in the group receiving the community standard of care is especially alarming. These results portend that the maltreated infants in this group are at exceedingly high risk of maladaptive outcomes on subsequent stage-salient issues of development. The typical intervention services provided to maltreated infants are not sufficient to remediate their early developmental difficulties. Without the receipt of theoretically derived, intensive interventions, maltreated infants are likely to embark on a negative developmental trajectory. Clearly, case management focused on monitoring the physical safety of maltreated children is inadequate to foster positive socioemotional development. Moreover, these results underscore the critical importance of providing evidence-based services to maltreating families to interrupt this negative cascade. The results of this randomized prevention trial are both heartening and sobering. The fact that plasticity is possible during infancy and that even the most disorganized form of attachment is modifiable in extremely dysfunctional mother-child dyads offers significant hope for thousands of young children and their families. By fostering secure attachment, costlier interventions, such as foster care placement, special education services, residential treatment, and incarceration, can be averted. Unfortunately, these results also shed light on the harsh reality of ineffective services currently being provided in communities throughout the nation. It is critical that professionals, government officials, social policy advocates, and mental health insurers recognize the necessity of investing in the delivery of theoretically informed, evidence-based interventions. Although this may appear to be a costly endeavor in the short term, the long-term benefits with respect to enhancing positive child development, preventing the dissolution of families, and decreasing the burden of mental illness in society cannot be overstated.

Randomized Control Trials Targeting Caregiving with Neurobiological-Level Evaluation

RCTs to promote child resilience by targeting the quality of caregiving also can measure effectiveness at biological levels of function in children. Research on the impact of early caregiving adversity in humans has revealed that the HPA axis is highly likely to exhibit evidence of dysregulation and that this dysregulation may entail both hyper-and hypofunctioning (Cicchetti & Rogosch, 2001b; Cicchetti, Rogosch, Gunnar, & Toth, 2010; Gunnar & Vazquez, 2001; Heim et al., 2008). Levels of cortisol, the final hormonal product of the neuroendocrine system, that are chronically elevated during development when neural circuits are still in the process of maturing exert an impact on brain development by affecting neurotropic factors and regulating gene transcription, thereby influencing the manner in which neural circuits perceive and interpret environmental threat and the magnitude and duration of future stress responses (Tarullo & Gunnar, 2006). In addition, chronically low levels of basal cortisol also can eventuate in harmful effects on the brain as well as on physical and mental health and behavior (Juster, McEwen, & Lupien, 2009; Juster et al., 2011; Lupien et al., 2006). Both hyper- and hypocortisolism can reflect allostatic load, typically defined as the result of chronic exposure to fluctuating or heightened neural or neuroendocrine responses that emerge in response to chronic environmental challenges that are perceived as especially stressful (McEwen & Stellar, 1993). Stressful experiences in early life are vital in the determination of the capacity for flexible/resilient adaptation or the lack thereof (McEwen, Gray, & Nasca, 2015). The plethora of studies documenting the neuroendocrine disturbances following child maltreatment, in concert with the likely harmful developmental sequelae of these HPA axis disturbances and the accompanying sustained allostatic load, have underscored the potential importance of RCTs to try to modify and reorganize alterations in HPA axis functioning (Cicchetti & Gunnar, 2008). In the human literature, evidence is beginning to accrue suggesting that efficacious interventions with children in foster care may modify not only maladaptive behavior but also the neuroendocrine correlates of behavior (Dozier et al., 2008; Fisher, Gunnar, Dozier, Bruce, & Pears, 2006; Fisher, Stoolmiller, Gunnar, & Burraston, 2007). Methodologically rigorous RCTs that incorporate a multilevel measurement battery have begun to provide a unique lens through which the processes responsible for the development, maintenance, and modification of both typical and atypical functional neuroendocrine outcomes can be discerned (Gunnar, Fisher, & the Early Experience, Stress, and Prevention Network, 2006). In an innovative illustration of the capability of psychosocial interventions to modify diurnal cortisol activity, noted briefly earlier, Fisher et al. (2007) conducted a 6- to 9-month RCT of a family-based foster care therapeutic intervention for preschool children. The therapeutic intervention was compared to a regular foster care condition and a demographically comparable group of same-aged, nonmaltreated preschool children from low-SES families. Fisher et al. (2007) collected and assayed early-morning and evening cortisol levels from the preschool children (mean age = 4.4 years) on a monthly basis from the baseline until the end of the interventions. One common disturbed cortisol pattern that emanates from disruptions in caregiving is a low early-morning cortisol that displays minimal change across the day. Fisher and colleagues found that the foster children who received the family-based therapeutic

intervention exhibited cortisol activity that became comparable to the nonmaltreated comparison children over the course of the intervention. In contrast, children who received routine services in state foster care (e.g., weekly individual psychotherapy; developmental screening for delays with the subsequent provision of special education services, if needed) displayed an increasingly flattened pattern of early-morning to evening cortisol activity throughout the course of the RCT. The results of this psychosocial intervention demonstrate that improvements in caregiving following the experiences of early adversity have the potential to reverse or prevent disturbances in the HPA system (Fisher et al., 2007). The study of child maltreatment as a “natural experiment” provides a valuable lens into the relations between sensitive caregiving and stress regulation and reveals insights for translation of basic empirical findings into RCTs aimed at reorganizing and normalizing these relations. Moreover, the conduct of RCTs, such as that already described, yields important knowledge regarding the malleability of developmental processes. Psychosocial randomized preventive interventions with previously maltreated children placed in foster care provide important opportunities for children who have transitioned into a changed environment to manifest competence in biological and behavioral functioning. For children in foster care, there may be variability with respect to the presence of continued stressful experiences. For example, the move to foster care may be traumatic for some children even in they have been placed in a better setting than their previous home or foster care experience. Longitudinal effects of child maltreatment on cortisol regulation in infants from 1 to 3 years were investigated in the context of an RCT (Cicchetti, Rogosch, Toth, & Sturge-Apple, 2011). Thirteen-month-old infants and their mothers were randomly assigned to one of three intervention conditions: child-parent psychotherapy (CPP), psychoeducational parenting intervention (PPI), and a control group involving standard community services (CS). A fourth group of infants from nonmaltreating families and their mothers comprised a nonmaltreated comparison (NC) group. The two active interventions were combined into one maltreated intervention (MI) group for statistical analyses. Saliva samples were obtained from children at 10:00 a.m. before beginning a laboratory observation session with their mothers when the children were 13 months of age (preintervention), 19 months (midintervention), 26 months (postintervention), and 38 months (1-year postintervention follow-up). At the initial assessment, no significant differences among groups in morning cortisol were observed. Latent growth curve analyses examined trajectories of cortisol regulation over time. Beginning at midintervention, divergence was found among the groups. Whereas the MI group remained indistinguishable from the NC group across time, the CS group progressively evinced lower levels of morning cortisol, statistically differing from the MI and NC groups. These results are consistent with those of Fisher et al. (2007) and Dozier et al. (2008) and highlight the value of psychosocial interventions for early child maltreatment in normalizing biological regulatory processes. Random assignment to adversity or trauma is not possible, but random assignment to intervention is. If biological systems recover in response to an RCT intervention, then this provides support for arguments that the systems under study are sensitive to environmental input during development. Furthermore, if randomized interventions alter neurobiological

systems associated with disorders, and it can be shown that they mediate positive changes in psychosocial functioning, then these intervention studies can foster a better understanding of the neurobiological contributors to resilience. The Bucharest Early Intervention Project (BEIP) is a landmark multilevel RCT preventive intervention informed by developmental theory (Nelson, Fox, & Zeanah, 2014). The project has three major goals: (1) to document the impact of institutionalization on child development across a number of salient domains; (2) to ascertain the extent to which children in foster care can recover from experiencing early adversity; and (3) to assist the Romanian government in building alternative means to institutionalization of caring for children experiencing adversity. BEIP is a foster care intervention for children who had been abandoned at or shortly after their birth and placed in an institution for young children in Bucharest, Romania (Fox et al., 2011; Nelson et al., 2007; Zeanah et al., 2003). Half of a group of 136 children residing in an institution in Bucharest were randomly assigned either to high-quality foster care designed specifically for the BEIP, and the other half were randomly assigned to remain in institutional care (care as usual group). A group of never-institutionalized children serves as community controls. Over time, the purity of this experiment was affected by naturally occurring changes in child status. However, it remains an extraordinarily rare effort to test the effects of foster care in comparison to institutional and typical family care. On average, children in the high-quality foster care group entered into foster care at 22 months and were assessed longitudinally at stage-salient developmental periods at 30, 42, 54, and 96 months. A follow-up assessment at age 12 years is under way, focused on early adolescence, which represents another critical developmental transition. A variety of psychological and biological domains were assessed at various developmental periods, with each transitional turning point thought to provide both opportunities and challenges for the developing child. Among the major findings of BEIP, it was demonstrated that children who were placed into foster care early in life showed improved cognitive, emotional, and behavioral functioning. In contrast, children randomly assigned to remain in institutionalized care showed ongoing developmental deficits across all domains assessed. Thus, early placement into foster care resulted in the most optimal outcomes for the formerly institutionalized children. Interestingly, children who were randomly placed in foster care demonstrated increases in cognitive functioning (IQ) and in brain development, whereas children who remained in institutions displayed lower levels of electrical activity across all brain regions (Vanderwert, Marshall, Nelson, Zeanah, & Fox, 2010). The inclusion of a normative community control group was an important addition to the BEIP design. It provided a more stringent test for evaluating the efficacy of the foster care interventions than would a foster care versus care as usual comparison. On several of the domains assessed in the BEIP, children randomly assigned to foster care performed similarly to children in the never-institutionalized group; however, for other domains, children in foster care were functioning in between the institutionalized (care as usual group) and neverinstitutionalized groups (e.g., on measures of IQ, emotional responsiveness and reactivity, attachment security, and anxiety disorders).

Miller, Brody, Yu, and Chen (2014) investigated whether a psychosocial intervention, focused on improving parenting, strengthening family relationships, and building youth competencies could reduce inflammation in low-SES, African American families residing in poverty. Families were randomly assigned to a 7-week psychosocial intervention or to a control condition. When youth reached age 19, peripheral blood was collected to quantify six cytokines that orchestrate inflammation, the dysregulation of which contributes to many of the health problems known to pattern by SES. Youth who participated in the intervention had significantly less inflammation on all six indicators relative to controls. Inflammation was lowest among youth who received more nuturant-involved parenting and less harshinconsistent parenting as a consequence of the intervention. These findings have theoretical implications for research on resilience to adversity and the early origins of disease. If substantiated, they may also highlight a strategy for practitioners and policymakers to use in ameliorating social and racial health disparities.

Moving toward RCTs in Multiple-Levels-of-Analysis Research on Resilience Developmental psychopathologists have advocated a multiple-levels-of-analysis approach to the investigation of maladaptation, psychopathology, and resilience (Cicchetti & Blender, 2006; Cicchetti & Curtis, 2006, 2007; Cicchetti & Dawson, 2002; Masten, 2007, 2014). Thus, the examination of multiple biological and psychological systems, domains, and levels of the ecology on each individual participant in an RCT preventive intervention should eventuate in a more complete understanding of individual patterns of maladaptation, psychopathology, and resilience. We think it is essential that this multiple-levels-of-analysis approach is adopted by prevention scientists who are conducting RCTs. For example, the inclusion of molecular genetic methods (e.g., sequencing DNA and utilizing functional polymorphisms to examine genetic and epigenetic methylation and moderation of intervention), neuroimaging techniques (e.g., to investigate brain structure, function, and connectivity, pre-and post-intervention), and stress-reactivity paradigms (e.g., to ascertain whether neurobiological stress systems are modifiable as a result of intervention), in conjunction with the incorporation of psychological/behavioral outcomes, will enhance the understanding of interventions on brainbehavior relations (Cicchetti & Blender, 2006; Fishbein, 2000; Zeanah et al., 2003). Similarly, the inclusion of biological measures into the research armamentaria of researchers investigating pathways to adaptive functioning in the presence of significant adversity, as well as the translation of this integrative research into resilience-promoting interventions that assess biological and psychological measures concurrently, will contribute to reducing the schisms that exist between the study of normal and abnormal development and between research and practice (Beauchaine, Nehaus, Brenner, & Gatzke-Kopp, 2008; Cicchetti & Toth, 2006; Masten, 2014b). The utilization of a neurobiological framework and the incorporation of genetically sensitive designs into interventions seeking to promote resilience or to repair positive adaptations gone awry may contribute to the ability to design individualized interventions that are based on knowledge gleaned from biological and psychological levels of analysis. If assessments of

biological systems are routinely incorporated into the psychological measurement batteries employed in resilience-promoting interventions, then researchers will be in a position to discover whether the nervous system and psychological processes have been modified by experience. A challenge that will need to be surmounted involves the determination of the mechanisms whereby different levels of analysis interact and transact across developmental time. Furthermore, researchers conducting their work at each level will need to develop theories that are consistent or coherent with respect to other key levels of analysis. Although investigations that focus solely on the behavioral level of resilience remain valuable, it is now important that researchers who examine the determinants of resilience integrate empirical work conducted on the behavioral level with the genetic and biological levels, study their coactions, and investigate the relations among these mutually influencing systems. Determining the multiple levels at which change is engendered through RCT prevention trials will provide more insights into the mechanisms of change, the extent to which neural plasticity may be promoted, and the interrelations between biological and psychological processes in maladaptation, psychopathology, and resilience (Cicchetti & Curtis, 2006). Moreover, preventive interventions with the most in-depth empirical support, based on integrative multilevel theories of normality, psychopathology, and resilience, can be implemented in effectiveness trials in community or real-world settings to reach the broadest number of people and prevent, alleviate, and ultimately cure suffering from mental disorders. Furthermore, the inclusion of biological assessments in evaluations of interventions designed to foster resilience will enable scientists to discover whether the various components of multifaceted interventions each exert a differential impact on separate brain systems. It is thus possible to conceptualize successful resilience-promoting interventions as examples of experience-dependent neural plasticity. Finally, although resilience is a dynamic developmental construct that, theoretically, is changing throughout the life course, it nonetheless is important to ascertain whether and, if so, when there are optimal timing windows for efforts to promote resilience. Specifically, it will be extremely important to know which resilience-promoting interventions are more effective when implemented in the early years of life and which interventions need to be timed later for optimal effect. Likewise, are such interventions more likely to achieve their goal of promoting resilience if they are instituted as closely as possible to the identification of the adverse event(s) or experience?

Genetic and Epigenetic Moderation of Intervention Despite rapid, promising advances, our understanding of genetic moderation of intervention outcome remains in its early stages. Most interventions have aimed to change the environment with little to no consideration for genetic involvement. Thus, interventions that consider both genes and environment are in a nascent state. In a meta-analysis of 22 RCTs, van IJzendoorn and Bakermans-Kranenburg (2015) found clear-cut experimental support for genetic differential susceptibility. A recent Special Section of the journal Development and Psychopathology, edited by Belsky and van IJzendoorn (2015), was devoted to examining genetic moderation of intervention efficacy. The focus of the Special Section was to test

differential susceptibility by examining the RCTs conducted to date on genetically moderated RCT interventions. To provide an example, Cicchetti, Toth, and Handley (2015) examined the genetic moderation of interpersonal psychotherapy (IPT) efficacy for economically disadvantaged women with MDD. Specifically, Cicchetti et al. examined whether genotypic variation in the corticotropin-releasing hormone receptor 1 gene (CRHR1) and the serotonin transporter gene (5-HTT) moderated effects of IPT on depressive symptoms over time. Cicchetti et al. also tested genotype moderation of IPT mechanisms social adjustment and perceived stress. A diverse sample of nontreatment-seeking urban women at or below the poverty level with infants was recruited from the community and randomized to individual IPT or enhanced community standard. Results revealed that changes in depressive symptoms over time depended on both intervention group and genotypes (5-HTTLPR and CRHR1). Moreover, multiple-group path analysis indicated that IPT improved depressive symptoms, increased social adjustment and decreased perceived stress at posttreatment among women with zero copies of the CRHR1 TAT haplotype only. Finally, improved social adjustment at postintervention significantly mediated the effect of IPT on reduced depressive symptoms at 8 months postintervention only for women with zero copies of the TAT haplotype. Post-hoc analyses of 5-HTTLPR were indicative of differential susceptibility, albeit among African American women only. Prevention scientists have recently begun to consider epigenetic mechanisms within the context of understanding the efficacy of intervention on developmental outcomes and resilient functioning. As noted earlier, epigenetic mechanisms may be a realistic target for intervention because of their reversibility. Epigenetically informed RCT preventive interventions can include methylation assays in their measurement batteries to evaluate the effects of the interventions on these mechanisms and to refine theory. Nothing can be done to prevent the DNA sequence variations that eventuate in the emergence of risky phenotypes. Yet DNA methylation changes in response to experiences such as child abuse and neglect could help to bring about the design, implementation, and multilevel prevention and intervention strategies that change expression of risky genes through reversal or demethylation and thereby promote healthy physical and mental health development. Szyf and Bick (2013) concluded that DNA methylation adaptations early in life are system-wide and that they involve multiple gene circuits. Methylation assays also can be conducted at the level of site-specific regions of candidate genes with known functional properties, such as the glucocorticoid receptor gene (Bick et al., 2012). Thus, because responses to social adversity such as child maltreatment are not limited to the brain and also likely affect peripheral tissues, it is also possible to assess the efficacy of an intervention in peripheral cells. Prevention and intervention strive to alter the environment in order to bring about positive outcomes. Research on epigenetics suggests that prevention and intervention also may change the epigenome and that this could result in improved outcomes. If researchers are to understand the processes thorough which early adverse experiences such as child maltreatment and institutionalization impact maladaptation, psychopathology, or resilience, then it is critical that genetic variation (functional polymorphisms) and epigenetic modifications are examined.

Research on GxE and on epigenetics needs to incorporate, as well as to emphasize, a developmental perspective (GxExD; Sroufe, 2013). Genes may influence how environmental experience affects development, and this moderating role may operate differently at various developmental periods. Moreover, the effects of genes and experience at a particular period may be influenced by the effects of prior development. Environments may affect the timing of genetic effects and gene expression. In addition, there are experience effects on the epigenome, and these effects also would operate differently across the course of development. Genetic effects on intervention efficacy can occur in a number of ways. Are some individuals, based on genetic variation, more susceptible to the positive effects of intervention? Are different interventions more effective with different individuals based on genetic differences (i.e., matching intervention to genotype group)? Does intervention affect DNA methylation, resulting in changes in gene expression? DNA methylation changes in response to experience could lead to the design of both prevention and intervention strategies that change the expression of genes to promote healthy physical and mental health outcomes. Given that the demethylated epigenome would be transmitted to the next generation, it will be important to determine if decreased child maltreatment risk through efficacious intervention alters the epigenome, which in turn results in a “less risky” epigenome being transmitted to offspring. The design and conduct of longitudinal multilevel RCTs that incorporate baseline preintervention, postintervention, and postintervention follow-up assessments of psychosocial, neurobiological, and epigenetic processes through the collection of DNA and RNA over developmental time represent an exciting direction for future intervention studies.

Implications of Resilience Science for Policy Resilience science has many implications for policymakers at many levels of governmental and nongovernmental organizations, and there is growing evidence that policymakers are taking notice, with surging popularity of the concept in initiatives and products of humanitarian and governmental agencies at local, state/provincial, national, and international levels (Ager, 2013). UNICEF (2011), for example, published the Humanitarian Action for Children: Building Resilience (2011) report. The World Bank published an edited volume on Children and Youth in Crisis: Protecting and Promoting Human Development in Times of Economic Shocks, with a strong emphasis on concepts and findings from a resilience perspective for mitigating the effects of economic shocks on child development (Lundberg & Wuermli, 2012). Educational policies and practices have adopted positive more models of practice, often drawing directly on resilience science (e.g., Cohen, 2013; Doll, 2013). The impact of a resilience perspective on child welfare reform was well documented in a conference and subsequent publication, Promoting Resilience in Child Welfare (Flynn, Dudding, & Barber, 2006). The global movement to shift child welfare toward a positive, strength-focused, and developmental approach was highlighted in the conference and book, which also underscored the appeal of this approach to the stakeholders involved, including young people as well as social workers and funders. Disaster preparedness and response approaches have also begun to incorporate resilience

perspectives at multiple levels, underscoring the interconnectedness of resilience at the child, family, and community levels, with important effects on policies (Institute of Medicine, 2014; Masten et al., 2015). Data from studies of resilience in children following disaster consistently indicated the importance (for minimizing harm and facilitating recovery) of keeping families together, restoring child care and school, and providing opportunities for children to play and return to other childhood and family activities, and these have become guiding priorities in disaster preparation and response policies. As noted, prevention science has incorporated many of the models and findings from resilience science, expanding the focus from preventing problems and interrupting negative progressions to promoting positive development, building assets, and facilitating competence cascades (Cicchetti, 2013; Gest & Davidson, 2011; Luthar & Cicchetti, 2000; Masten, 2014b; Zimmerman et al., 2013). Universal prevention programs, as well as targeted programs for high-risk families or children, now often emphasize building resources in the family, school, or community (or all three) and competence skills in individual children, parents, or teachers (National Research Council and Institute of Medicine, 2009; Winslow, Sandler, Wolchik, & Carr, 2013). There appears to be a broad consensus among many stakeholders involved in policy that resilience approaches are valuable. The evidence base grows stronger for efficacy of interventions in many domains of adaptive development and for different situations that can be harmful to children. In addition, parents and policymakers alike find models appealing that focus on promoting positive development and achieving success (Ager, 2013; National Research Council and Institute of Medicine, 2009; Sandler et al., 2011). At the same time, evidence is building that well-timed and targeted intervention can have high long-term returns on public investments, benefiting the individual children or families as well as society (e.g., Cunha & Heckman, 2011; Heckman, 2006; Reynolds, Temple, White, Ou, & Robertson, 2011). As resilience science moved toward dynamic models of interacting systems, the implications for policy have concomitantly shifted to encompass more systems and the idea of integrated interventions that address more than one system at a time in order to generate synergy (Masten, 2011, 2014b, 2014c). Policymakers often must choose where and when to invest most strategically to promote the well-being of their constituents in the near and distant future. Research on the economics of human capital investment, naturally occurring resilience, and prevention science is aligning in helpful ways for applications in policy, although it still remains crucial to communicate the evidence and its implications effectively (Winslow et al., 2013). At the level of global intervention and investment in children, momentum is growing to raise the bar of humanitarian aid beyond child survival to children with the capacity to thrive, even in a world still plagued by war, epidemics, and extreme poverty (Masten, 2014a). Early in 2014, the U.S. National Academies launched a new international forum, “Investing in Young Children Globally,” to assemble evidence about best practices worldwide for strategic investments in children to promote lifelong health and well-being in individuals and societies (see their website at

http://iom.edu/Activities/Children/InvestingYoungChildrenGlobally.aspx). This international network of scientists, clinicians, policymakers, and other professionals convenes workshops and other meetings around the world to evaluate, integrate, and disseminate promising data and strategies for governments and nongovernmental organizations to invest effectively in young children in order to promote the healthy development of children and their societies. This effort produces workshop reports, infographics, and other products for widespread distribution to policymakers and funders worldwide. Also in 2014, a consortium of partners from multiple countries initiated a new set of three initiatives as part of the ongoing global “Grand Challenges” effort launched originally a decade earlier by the Bill and Melinda Gates Foundation, the Canadian government, and others. “All Children Thriving” is one of the new Grand Challenges initiatives with the goal of helping mothers and children thrive in the developing world.

Controversies Old and New Beginning with the first generation of resilience science, controversies emerged (Masten, 2012, 2014b; Rutter, 1987). Some of these controversies have waned as the science matured, some have returned in new forms, and some new issues have surfaced. All these trends are expectable in a maturing science and actually may reflect real progress. Here we highlight several of the enduring controversies and issues for research on resilience in human development.

Theoretical Issues From early in the study of child resilience, there were debates about how to define the concept, and these issues continue to be the subject of debate (Masten, 2014b; Panter-Brick & Leckman, 2013; Southwick, Bonanno, Masten, Panter-Brick, & Yehuda, 2014). Although there is general agreement that resilience science is focused on understanding adaptation in contexts of challenge or adversity, definitions of resilience vary, which leads to issues in operationalizing the concept in research (Luthar, 2006; Masten 2014b). Some definitions continue to focus on a traitlike concept of resilience (see later discussion) while others focus on capacities for, processes of, or outcomes of positive adaptation in the wake of adversity or contexts or risk. Even for those who accept a systems definition, where resilience refers to capacity or processes of a system engaged in adapting in response to a significant challenge, there are issues about focusing on internal function of the system or its adaptation to the environment. These issues reflect the nature of adaptation in living systems, which are always functioning as a unit embedded and interacting with other systems. Methodologically, this situation leads some investigators to focus on psychological well-being or internal indicators of function, such as allostatic load, while others focus on competence in the environment (e.g., social and academic competence) and still others include both kinds of adjustment indicators. Striking consistencies in findings have emerged in the resilience literature despite the variations in concepts and methods (Masten, 2014b). It is nonetheless problematic that the

variations in conceptual approaches make it difficult to integrate theory and findings on resilience. Greater effort to harmonize research concepts, designs, and measures would greatly facilitate gains in knowledge in resilience science.

Empirical Issues Numerous methodological issues flow from the conceptual variations in resilience science and also from the inherent complexity of research on dynamic living systems. Any competent and reflective researcher engaged in studies of resilience must confront these issues and make choices about concepts, models, measures, and methods. One of the major issues confronting resilience researchers is how to define and measure the core criteria of “risk” and “adaptation” delineated in the theoretical section of this chapter. Some investigators study a single risk marker, such as “homeless” or “premature” or “maltreated,” fully cognizant that these indicators represent a cumulative index of risks and resources. Other investigators deliberately assess cumulative risk or adversity indicators, and they must make choices about how to aggregate them for purposes of analysis (Evans et al., 2013; Obradović et al., 2012). These may include counts or other kinds of composite scores on sociodemographic indicators, perceived stress, or adverse experiences. Similarly, choosing what to assess on the adaptation side of the equation also is challenging. Some investigators focus on success in age-salient developmental tasks while others focus on symptoms of mental illness or psychological distress. Investigators are more aware now of the cultural influences on these choices, which often were ignored in early studies (Masten, 2014b; Panter-Brick & Leckman, 2013; Ungar et al., 2013). In international research, much more attention is given now to the issue of defining and measuring resilience in culturally meaningful ways. Many choices are made by investigators when they define resilience and choose measures of risk/adversity, adaptive function or outcome, and potential promotive or moderating influences. These choices have yielded an incredibly diverse and wide-ranging set of findings and also posed challenges for empirical methods of compiling the findings, such as meta-analysis. Again, coordinated studies could be extremely valuable in future resilience research.

Is Resilience a Trait? One of the most tenacious and wrong-headed issues in resilience science is the idea that resilience is a trait (Masten, 2012, 2014b; Panter-Brick & Leckman, 2013; Rutter, 1979). This enduring idea may stem from the origins of the word resilience itself, in that it often refers to a property of a thing or substance—the resiliency of a spring or a rubber band for instance. Early concepts of “hardiness” (Kobasa, 1979) and “ego resiliency” (Block & Block, 1980) in personality theory also contributed to this notion. While traits have been implicated in the observed capacity of human individuals to withstand or recover from major blows or challenges, there is no single trait that could be construed as “resilience.” The idea that resilience is a trait ignores overwhelming evidence that the

adaptation of individuals to adversity depends on relationships and many other systems external to the individual. Moreover, this is a dangerous idea because it implies that children or adults who struggle to adapt or have not yet recovered from traumatic experiences are somehow lacking the “right stuff,” in effect blaming the victim when the causes of stress or breakdown have much more to do with overwhelming adversity exposure and lack of socioecological supports or protections. Furthermore, there is ever-growing evidence that the same trait that functions in one set of conditions as a protective influence on function or development can function as a vulnerability in another context or at another point of time in development. Early examples identified by resilience investigators included individual differences in temperament as well as the biological example of sickle-cell trait (Rutter, 1987; Masten, 2012, 2014b). Recent research on differential susceptibility or biological sensitivity to context described previously in this chapter underscore the point that a trait may have different roles in different contexts.

Do Genes Determine Resilience? The trait debate has returned in a new guise with the research on genotypes as singular promotive or protective factors. This deterministic view is completely unreasonable and simplistic, given that all human function and development is influenced by many interactions of genes with context and many forms of experience. The capacity of any individual for adaptation is highly complex, develops over time as a result of many interactions of the individual with the environment, and at any given time is distributed across multiple systems, including social systems external to the person.

Is Adversity Good for Children? The idea that challenges can promote growth or, to paraphrase the famous words of Nietzsche quoted earlier that “whatever does not kill you makes you stronger,” is another concept that keeps returning as an issue in resilience science. As discussed in the section on steeling effects and posttraumatic growth, there is some evidence suggesting that modest challenges and the recovery that follows can strengthen capabilities involved in resilience. This idea was captured by the challenge model of resilience early in the science (Garmezy et al., 1984) as well as in the notion of steeling effects (Rutter, 1987). Contemporary data on adaptive systems (e.g., the immune or stress-reactivity systems) are congruent with the idea that some challenge is helpful for adaptive systems to organize and calibrate themselves effectively for the expected environment. However, we want to emphasize again that there is no evidence that severe adversity and trauma is good for development. As in the situation of surviving Ebola as a method for achieving immunity, the risks and costs of severe threats to development, exemplified by child abuse and other forms of child exploitation or suffering, far exceed any conceivable benefit. When severe adversity does occurs, though, certainly it is the case that support and recovery are better for development than continued adversity or neglect.

Is There a Cost to Resilience?

A final example of enduring issues in resilience science is raised by the intriguing question of whether there is a price to pay for resilience (Masten, 2014b, Masten & Coatsworth, 1998). A number of influential studies raised the possibility that the stress or strain of overcoming adversity could cause current distress or later health issues (Luthar, 1991; Luthar, Doernberger, & Zigler, 1993; Werner & Smith, 2001). Some of these “costs” may be attributable to the adversity itself, akin to scars from battle, whereas others may result from the strains of setting high goals and related adaptive efforts. Recovering child soldiers and survivors of severe maltreatment, for example, would not be expected to come away from terrible experiences with no residual effects, but these effects may be better described as the costs of chronic trauma rather than resilience. This question has resurfaced in recent work on allostatic load. In a provocatively titled article (“Is Resilience Only Skin Deep?”), Brody and colleagues (2013) reported that some of the most successful young people in their study of high-risk African American youth were showing signs of high allostatic load despite their positive psychosocial adaptation. They raised the possibility that overcoming the risks of economic strains and racism may take a toll on the body in the form of chronic wear and tear, indexed by high blood pressure or body mass index. As they follow these young people, it will be important to track their adaptation at multiple levels of analysis to see if success at one level continues to have costs at another level.

Progress and Future Directions Progress and transformation in resilience science over the past decade is marked in many ways by implementation of a dynamic, multiple-levels perspective in conceptual, methodological, and empirical gains. This is the hallmark of the fourth wave, and it has set the stage for the coming decade.

Advances in Models and Methods In conceptual models, the shift toward dynamic, transactional theory transformed resilience science to reflect the growing prevalence of relational developmental systems theory as the core perspective in developmental science (e.g., Overton, 2013, Zelazo, 2013). Definitions of resilience, although still sometimes criticized and controversial, dramatically tilted toward systems-based definitions that are compatible with related concepts in other sciences that focus on other system-levels, including family resilience, community resilience, and ecological resilience. With this shift, the idea of resilience as a trait lost support, and interest grew in the processes linking multiple systems to influence development. Among the most prominent conceptual drivers of theory and research on resilience in the past decade were GxE concepts and particularly the roles of epigenetic processes in adaptation that could influence resilience (Cicchetti & Rogosch, 2012; Karatsoreos & McEwen, 2013; KimCohen & Gold, 2009; Kim-Cohen & Turkewitz, 2012). Conceptually, these ideas renewed theoretical and empirical interest in the programming of adaptive systems important for resilience (as well as vulnerabilities) and also sparked new thinking about intervention.

The concept of individual differences in the openness of an individual to environmental influence, combined with recognition of developmental windows of greater or lesser plasticity and sensitive periods for adaptive tuning of the individual to the expected environment, is transforming theory and research on the roles of both genes and experience in resilience (Del Giudice et al., 2011; Ellis et al., 2011; Masten, 2014b). These ideas and the empirical studies they inspire remain a key frontier for resilience research in the coming decade, particularly for intervention experiments designed to promote resilience through strategic timing and targeting or to program/reprogram adaptive systems to increase resilience. The concept of developmental cascades, where function at one system level can spread to affect other levels through intersystem interactions, attracted considerable interest, including two special issues of the journal Development and Psychopathology (Masten & Cicchetti, 2010a, 2010b). These cascades can be bidirectional with neurobiological changes influencing psychosocial processes or vice versa. In addition, there was growing interest in delineating and testing models of how resilience in contextual systems of a child's life, such as a family or school, spreads to boost the resilience of individual children within the family or school. These resilience cascades can be naturally occurring or result from deliberate efforts to induce a cross-system cascade, as in interventions to foster better parenting as a means to normalize the biological stress regulation system (e.g., Fisher et al., 2007). Empirically, we would expect research on these intersystem, multiple-level cascades to expand. Methodological advances also reflect improved capabilities in the science to measure and analyze multilevel and multivariate models (Masten, 2014b). For example, testing for developmental cascade effects requires repeated measures over time and statistical approaches that can account for inevitably missing data and multiconcept, transactional models. In addition, practical advances in structural equation techniques and software for analyzing interindividual growth and linking such growth to other variables or growth parameters at other levels also became feasible. Tools for assessing neurobiological function and change also spurred research on processes that might account for resilience (Masten, 2014b). Perhaps most dramatic was the rapid increase in genetic tools for assaying DNA and epigenetic change at a more reasonable cost. Additionally, new tools for assays of cortisol in hair as well as saliva were helpful in research on stress and adaptation, and strategies for safely and meaningfully imaging brain function in very young children also advanced. We expect these tools to be put to even more use in resilience science going forward.

Advances in the Knowledge Base Research on resilience has expanded dramatically at the biological level of focus, as noted. After many calls for greater attention to the neurobiology of resilience, numerous studies were undertaken with many more doubtless in progress. Psychosocial findings on resilience are now being studied at multiple levels to shed light on how change occurs, through both biological and social processes and intersystem interactions, and we expect the knowledge base to expand.

At the same time, there also has been remarkable progress in conceptual and empirical work on resilience in sociocultural context, in two major ways. First, there is simply more research in more diverse cultures focused on resilience, ranging from studies of resilience in child soldiers of Africa to survivors of disasters in regions all around the world, from China to Chile and Finland to South Africa. There is particularly more research on resilience in children of low-and middle-income countries, long neglected heretofore. Second, there is more research focused on cultural similarities and differences in resilience processes. Specific areas of expanding science on resilience in cultural contexts include multinational studies, research on immigrant youth, research on investing in children globally, and studies of children in ongoing political conflicts and war, including refugees (Masten, 2014a, 2014b). Research from diverse cultures supports to a surprising degree many of the major findings from earlier generations of resilience studies limited to a few cultures or economically developed countries. However, culturally focused research also highlights cultural differences in resilience and the culturally-grounded practices of communities designed to promote resilience (Ungar et al., 2013; Ungar, 2012). This decade witnessed many additional and higher-quality intervention experiments pertinent to resilience, often in the form of gold standard RCTs (Cicchetti et al., 2006; Cicchetti et al., 2015; Masten, 2014b; Masten et al., 2015). Examples have provided throughout this chapter of intervention trials, including examples that illustrate how RCTs are beginning to include multiple levels of assessment that make it possible to study GxE and GxI effects and multiplelevel cascades. We expect these directions of research, with their triple benefits—to theory, practice, and policy—to continue and expand in the coming decade. More broadly, there is a surge of resilience-oriented research focused on investing in children and particularly children at risk globally, in both developing and developed nations (Lundberg & Wuermli, 2012; Masten, 2014a). Investment models for international organizations like UNICEF and the World Bank increasingly appear to reflect a focus on promoting competence and resilience, moving beyond survival. Additionally, international aid organizations are seeking evidence-based practices that build human capital and integrate resources across system levels to promote positive child development in very disadvantaged populations. The past decade also is characterized by increasing attention to resilience in policies aimed at the well-being of children and adults at the local, national, and international level (Ager, 2013; Masten, 2014a; Panter-Brick & Leckman, 2013). We also see this translational application of resilience science continuing and also hope that it will be guided by the best possible research.

Future Directions Given the progress and transformations we have documented in this chapter, we anticipate a fruitful decade of resilience science to come. We expect a continuing focus on multiple-level processes in models and methods that will fill in the many blanks in our understanding of how resilience works within and across levels, from the molecular level to the levels of individuals, families, schools, communities, cultures, media, and global policy.

As knowledge and tools continue to improve on the human genome and epigenetic processes, we expect a surge in theory-building and studies of resilience processes, particularly those that explain how naturally- and experimentally-induced adaptation influence resilience. We also expect more research on effective timing and targeting of interventions with respect to individual and cultural differences, biological programming, developmental cascades, and recovery pathways. We anticipate that there will be growing attention to the synergistic effects of integrated interventions that address multiple processes simultaneously to change the odds in favor of resilience and positive development. In the coming decade, we also hope to see evidence that adaptive systems that developed in less optimal ways due to high adversity and inadequate protections can be recalibrated by new experiences. It will be exciting to see interventions that build resilience in the lives of children whose development may be compromised by residual effects of adversity or lack of protection. Finally, we expect that the developmental study of resilience over the life span will become more integrated. A number of longitudinal studies are following cohorts later into adulthood, and the research on resilience in late life is expanding. Long-term studies of resilience have the potential to illuminate the benefits and costs of both adversity and resilience for future health and well-being. The future of resilience science at this juncture appears to be bright with promise. No doubt critical evaluations of the theory and evidence will continue, as expected in a mature science. Challenges and effective responses to such challenges can build a stronger science. Thus far, resilience science appears to be robust in the face of challenges and contributing in meaningful ways to practice and policy that benefits children and societies, just as the pioneers envisioned.

References Adler, N. E., & Ostrove, J. M. (2006). Socioeconomic status and health: What we know and what we don't. Annals of the New York Academy of Sciences, 896, 3–15. doi: 10.1111/j.1749– 6632.1999.tb08101.x Adolphs, R. (2009). The social brain: Neural basis of social knowledge. Annual Review of Psychology, 60, 693–716. doi: 10.1146/annurev.psych.60.110707.163514 Ager, A. (2013). Annual research review: Resilience and child wellbeing—public policy implications. Journal of Child Psychology and Psychiatry, 54, 488–500. doi: 10.1111/jcpp.12030 Ainsworth, M. D. S. (1989). Attachments beyond infancy. American Psychologist, 44, 709– 716. doi: 10.1037/0003–066X.44.4.709 Ainsworth, M. D., Blehar, M. C., Walters, E., & Wall, S. (1978). Patterns of attachment: A psychological study of the Strange Situation. Hillsdale, NJ: Erlbaum.

Akos, P., & Galassi, J. P. (2008). Strengths-based school counseling: Introduction to the special issue. Professional School Counseling, 12, 66–67. doi: 10.5330/PSC.n.2010–12.66 Alexander, G. R., & Kotelchuck, M. (2001). Assessing the role and effectiveness of prenatal care: History, challenges, and directions for future research. Public Health Reports, 116, 306– 316. doi: 10.1093/phr/116.4.306 Alink, L. R. A., Cicchetti, D., Kim, J., & Rogosch, F. A. (2009). Mediating and moderating processes in the relation between maltreatment and psychopathology: Mother–child relationship quality and emotion regulation. Journal of Abnormal Child Psychology, 37, 831– 843. doi: 10.1007/s10802–009–9314–4 American Psychological Association. (2010). Resilience and recovery after war: Refugee children and families in the United States. Washington, DC: Author. Anderson, A. R., Christenson, S. L., Sinclair, M. F., & Lehr, C. A. (2004). Check & Connect: The importance of relationships for promoting engagement with school. Journal of School Psychology, 42, 95–113. doi: 10.1016/j.jsp.2004.01.002 Anthony, E. K., Alter, C. F., & Jenson, J. M. (2009). Development of a risk and resilience-b ased out-of-school time program for children and youths. Social Work, 54, 45–55. doi: 10.1093/sw/54.1.45 Anthony, E. J., & Koupernik, C. (Eds.). (1974). The child in his family: Children at psychiatric risk (Vol. 3). New York, NY: Wiley. Arnold, S. E. (1999). Neurodevelopment abnormalities in schizophrenia: Insights from neuropathology. Development and Psychopathology, 11, 439–456. doi: 10.1017/S095457949900214X Baker, D. P., Salinas, D., & Eslinger, P. J. (2012). An envisioned bridge: Schooling as a neurocognitive developmental institution. Developmental Cognitive Neuroscience, 25, S6– S17. doi: 10.1016/j.dcn.2011.12.001 Bakermans-Kranenburg, M. J. & van IJzendoorn, M. H. (2008). Oxytocin receptor (OXTR) and serotonin transporter (5-HTT) genes associated with observed parenting. Social Cognitive and Affective Neuroscience, 3, 128–134. Bandura, A. (1997). Self-efficacy: The exercise of control. New York, NY: Freeman. Barber, B. K. (2008). Contrasting portraits of war: Youths' varied experiences with political violence in Bosnia and Palestine. International Journal of Behavioral Development, 32, 298–309. doi: 10.1177/0165025408090972 Barber, B. K. (Ed.). (2009). Adolescents and war: How youth deal with political violence. New York, NY: Oxford University Press. Barber, B. K. (2013). Annual research review: The experience of youth with political

conflict–challenging notions of resilience and encouraging research refinement. Journal of Child Psychology and Psychiatry, 54, 461–473. doi: 10.1111/jcpp.12056 Beauchaine, T. P., Nehaus, E., Brenner, S. L., & Gatzke-Kopp, L. (2008). Ten good reasons to consider biological processes in prevention and intervention research. Development and Psychopathology, 20, 745–774. doi: 10.1017/S0954579408000369 Becvar, D. S. (Ed.). (2013). Handbook of family resilience. New York, NY: Springer. doi: 10.1007/978–1–4614–3917–2 Beeghly, M., & Tronick, E. (2011). Early resilience in the context of parent–infant relationships: A social developmental perspective. Current Problems in Pediatric and Adolescent Health Care, 41, 197–201. doi: 10.1016/j.cppeds.2011.02.005 Beelmann, A., & Heinemann, K. S. (2014). Preventing prejudice and improving intergroup attitudes: A meta-analysis of child and adolescent training programs. Journal of Applied Developmental Psychology, 35, 10–24. doi: 10.1016/j.appdev.2013.11.002 Belsky, J., Bakermans-Kranenburg, J. M., & van IJzendoorn, M. H. (2007). For better and for worse: Differential susceptibility to environmental influences. Current Directions in Psychological Science, 16, 300–304. doi: 10.1111/j.1467–8721.2007.00525.x Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135, 885–908. doi: 10.1037/a0017376 Belsky, J., & van IJzendoorn, M. H. (Eds.) (2015). What works for whom?: Genetic moderation of intervention efficacy. [Special issue]. Development and Psychopathology, 27, 1–318. doi: 10.1017/S0954579414001254 Betancourt, T. S., Borisova, I. I., Williams, T. P., Brennan, R. T., Whitfield, T. H., de la Soudiere, M…Gilman, S. E. (2010). Sierra Leone's former child soldiers: A follow-up study of psychosocial adjustment and community reintegration. Child Development, 81, 1076–1094. doi: 10.1111/j.1467–8624.2010.01455.x Betancourt, T. S., McBain, R., Newnham, E. A., & Brennan, R. T. (2013). Trajectories of internalizing problems in war-affected Sierra Leonean youth: Examining conflict and postconflict factors. Child Development, 84, 455–470. doi: 10.1111/j.1467– 8624.2012.01861.x Bick, J., Naumova, O., Hunter, S., Barbot, B., Lee, M., Luthar, S. S.,…Grigorenko, E. L. (2012). Childhood adversity and DNA methylation of genes involved in the hypothalamus– pituitary–adrenal axis and immune system: Whole-genome and candidate-gene associations. Development and Psychopathology, 24(4), 1417–1425. doi: 10.1017/S0954579412000806 Bierman, K. L. (2011). The promise and potential of studying the “invisible hand” of teacher influence on peer relations and student outcomes: A commentary. Journal of Applied Developmental Psychology, 32, 297–303. doi: 10.1016/j.appdev.2011.04.004

Bierman, K. L., Domitrovich, C. E., Nix, R. L., Gest, S. D., Welsh, J. A., Greenberg, M. T.,… Gill, S. (2008). Promoting academic and social–emotional school readiness: The Head Start REDI program. Child Development, 79, 1802–1817. doi: 10.1111/j.1467–8624.2008.01227.x Bierman, K. L., Nix, R. L., Greenberg, M. T., Blair, C., & Domitrovich, C. E. (2008). Exeuctive funtions and school readiness intervention: Impact, moderation, and mediation in the Head Start REDI program. Development and Psychopathology, 20, 821–843. doi: 10.1017/S0954579408000394 Black, J. E., Jones, T. A., Nelson, C. A., & Greenough, W. T. (1998). Neuronal plasticity and the developing brain. In J. D. Noshpitz, N. E. Alessi, J. T. Coyle, S. I. Harrison, and S. Eth (Eds.), Handbook of child and adolescent psychiatry. Vol. 6: Basic psychiatric science and treatment (pp. 31–53). New York, NY: Wiley. Blair, C., & Raver, C. C. (2012). Individual development and evolution: Experiential canalization of self-regulation. Developmental Psychology, 48, 647–657. doi: 10.1037/a0026472 Block, J. H., & Block, J. (1980). The role of ego-control and ego-resiliency in the organization of behavior. In W. A. Collins (Ed.), Development of cognition, affect, and social relations. Minnesota Symposia on Child Psychology, 13, 39–101. Bonanno, G. A. (2004). Loss, trauma, and human resilience: have we underestimated the human capacity to thrive after extremely aversive events? American Psychologist, 59, 20–28. doi: 10.1037/0003–066X.59.1.20 Bonanno, G. A., Colak, D. M., Keltner, D., Shiota, L., Papa, A., Noll, J. G.,…Trickett, P. K. (2007). Context matters: The benefits and costs of expressing positive emotion among survivors of childhood sexual abuse. Emotion, 7, 824–837. doi: doi: 10.1037/1528– 3542.7.4.824 Bonanno, G. A., & Diminich, E. D. (2013). Annual Research Review: Positive adjustment to adversity–trajectories of minimal–impact resilience and emergent resilience. Journal of Child Psychology and Psychiatry, 54(4), 378–401. doi: 10.1111/jcpp.12021 Bonanno, G. A., Westphal, M., & Mancini, A. D. (2011). Resilience to loss and potential trauma. Annual Review of Clinical Psychology, 7, 511–535. doi: 10.1146/annurev-clinpsy032210–104526 Botvin, G. J., & Griffin, K. W. (2004). Life skills training: Empirical findings and future directions. Journal of Primary Prevention, 25, 211–232. doi: 10.1023/B:JOPP.0000042391.58573.5b Botvin, G. J., & Tortu, S. (1988). Preventing adolescent substance abuse through life skills training. In R. H. Price, E. L. Cowen, R. P. Lorion, & J. Ramos McKay (Eds.), Fourteen ounces of prevention: A casebook for practitioners (pp. 98–110). Washington, DC: American

Psychological Association. Bowes, L., & Jaffee, S. R. (2013) Biology, genes and resilience: Towards a multidisciplinary approach. Trauma, Violence & Abuse, 14(3), 195–208. doi: 10.1177/1524838013487807 Bowlby, J. (1982). Attachment and loss. New York, NY: Basic Books. (Original work published 1969.) Boyce, W. T., & Ellis, B. J. (2005). Biological sensitivity to context: I. An evolutionary– developmental theory of the origins and functions of stress reactivity. Development and Psychopathology, 17, 271–301. doi: doi: 10.1017/S0954579405050145 Bradley, B., Westen, D., Mercer, K. B., Binder, E. B., Jovanovic, T., Crain, D.,…Heim, C. (2011). Association between childhood maltreatment and adult emotional dysregulation in a low-income, urban, African American sample: Moderation by oxytocin receptor gene. Development and Psychopathology, 23(2), 439–452. doi: 10.1017/S0954579411000162 Bretherton, I., & Munholland, K. A. (1999). Internal working models in attachment relationships: A construct revisited. In J. Cassidy & P. R. Shaver (Eds.), Handbook of attachment: Theory, research, and clinical applications (pp. 89–111). New York, NY: Guilford Press. Britto, P. R., Engle, P. L., & Super, C. M. (Eds.). (2013). Handbook of early childhood development research and its impact on global policy. New York, NY: Oxford University Press. doi: 10.1093/acprof:oso/9780199922994.001.0001 Brody, G. H., Beach, S. R. H., Philibert, R. A., Chen, Y., & Murry, V. M. (2009). Prevention effects moderate the association of 5-HTTLPR and youth risk behavior initiation: Gene × environment hypotheses tested via a randomized prevention design. Child Development, 80, 645–661. doi: 10.1111/j.1467–8624.2009.01288.x Brody, G. H., Chen, Y. F., & Beach, S. R. (2013). Differential susceptibility to prevention: GABAergic, dopaminergic, and multilocus effects. Journal of Child Psychology and Psychiatry, 54, 863–871. doi: 10.1111/jcpp.12042 Brody, G. H., Murry, V. M., Gerrard, M., Gibbons, F. X., Molgaard, V., McNair, L., et al. (2004). The Strong African American Families Program: Translating research into prevention programming. Child Development, 75, 900–917. doi: 10.1111/j.1467–8624.2004.00713.x Brody, G. H., Murry, V. M., Kogan, S. M., Gerrard, M., Gibbons, F. X., Molgaard, V.,…Chen, Y.-F. (2006). The Strong African American Families Program: A cluster-randomized prevention trial of long-term effects and a mediational model. Journal of Consulting and Clinical Psychology, 74, 356–366. doi: 10.1037/0022–006X.74.2.356 Bronfenbrenner, U. (1979). The ecology of human development: Experiments by nature and design. Cambridge, MA: Harvard University Press.

Bronfenbrenner, U., & Morris, P. A. (1998). The ecology of developmental processes. In W. Damon & R. M. Lerner (Eds.), Handbook of child psychology: Vol. 1. Theoretical models of human development (5th ed., pp. 993–1028). New York, NY: Wiley. Brown, G. W., & Harris, T. (1978). Social origins of depression: A study of psychiatric disorder in women. New York, NY: Free Press. Burt, K. B., Coatsworth, J. D., & Masten, A. S. (2016). Competence and psychopathology in development. In D. Cicchetti (Ed.), Developmental Psychopathology (3rd ed., Vol. IV, pp. 435– 484). New York: Wiley. Calhoun, R. G., & Tedeschi, R. G. (Ed.). (2006). Handbook of posttraumatic growth: Research & practice. Mahwah, NJ: Erlbaum. Carlson, S. M., Zelazo, P. D., & Faja, S. (2013). Executive function. In P. D. Zelazo (Ed.), Oxford handbook of developmental psychology (pp. 706–743). New York, NY: Oxford University Press. doi: 10.1093/oxfordhb/9780199958450.013.0025 Carlson, V., Cicchetti, D., Barnett, D., & Braunwald, K. (1989). Disorganized/disoriented attachment relationships in maltreated infants. Developmental Psychology, 25, 525–531. doi: 10.1037/0012–1649.25.4.525 Carter, C. S., & Porges, S. W. (2013). The biochemistry of love: An oxytocin hypothesis. EMBO Reports, 14, 12–16. doi: 10.1038/embor.2012.191 Carter, C. S., & Porges, S. W. (2014). Peptide pathways to peace. In J. F. Leckman, C. PanterBrick, & R. Salah (Eds.), Pathways to peace: The transformative power of families and child development (pp. 43–64). Cambridge MA: MIT Press. Caspi, A., Hariri, A., Holmes, A., Uher, R., & Moffitt, T. E. (2010). Genetic sensitivity to the environment: The case of the serotonin transporter gene (5-HTT) and its implications for studying complex diseases and traits. American Journal of Psychiatry, 167, 509–527. doi: 10.1176/appi.ajp.2010.09101452 Caspi, A., McClay, J., Moffitt, T., Mill, J., Martin, J., Craig, I. W.,…Poulton, R. (2002). Role of genotype in the cycle of violence in maltreated children. Science, 297, 851–854. doi: 10.1126/science.1072290 Caspi, A., Sugden, K., Moffit, T. E., Taylor, A., Craig, I. W., & Harrington, H. L. (2003). Influence of life stress on depression: Moderation by a polymorphism in the 5-HTT gene. Science, 301, 386–389. doi: 10.1126/science.1083968 Cassidy, J., & Shaver, P. (Eds.). (2008). Handbook of attachment: Theory, research, and clinical applications. New York, NY: Guilford Press. Catani, C., Kohiladevy, M., Ruf, M., Schauer, E., Elbert, T., & Neuner, F. (2009). Treating children traumatized by war and tsunami: A comparison between exposure therapy and

meditation-relaxation in North-East Sri Lanka. BMC Psychiatry, 9, 22. doi: 10.1186/1471– 244X-9–22 Charney, D. (2004). Psychobiological mechanisms of resilience and vulnerability: Implications for successful adaptation to extreme stress. American Journal of Psychiatry, 161, 195–216. doi: 10.1176/appi.ajp.161.2.195 Christenson, S. L., Thurlow, M. L., Sinclair, M. F., Lehr, C. A., Kaibel, C. M., Reschly, A. L., Mavis, A. L., & Pohl, A. (2008). Check and connect: A comprehensive student engagement intervention manual. Minneapolis, MN: University of Minnesota, Institute on Community Integration. Ciaranello, R., Aimi, J., Dean, R. S., Morilak, D., Porteus, M. H., & Cicchetti, D. (1995). Fundamentals of molecular neurobiology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Vol. 1 Theory and method (pp. 109–160). New York, NY: Wiley. Cicchetti, D. (1984). The emergence of developmental psychopathology. Child Development, 55, 1–7. doi: 10.2307/1129830 Cicchetti, D. (1990). A historical perspective on the discipline of developmental psychopathology. In J. Rolf, A. Masten, D. Cicchetti, K. Nuechterlein, & S. Weintraub (Eds.), Risk and protective factors in the development of psychopathology (pp. 2–28). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511752872.003 Cicchetti, D. (1993). Developmental psychopathology: Reactions, reflections, projections. Developmental Review, 13, 471–502. doi: 10.1006/drev.1993.1021 Cicchetti, D. (2002). How a child builds a brain: Insights from normality and psychopathology. In W. Hartup & R. Weinberg (Eds.), Minnesota symposia on child psychology: Child psychology in retrospect and prospect (Vol. 32, pp. 23–71). Mahwah, NJ: Erlbaum. Cicchetti, D. (2006). Development and psychopathology. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology: Vol. 1. Theory and method (2nd ed., pp. 1–23). Hoboken, NJ: Wiley. Cicchetti, D. (Ed.). (2007). G×E interactions and developmental psychopathology [Special issue]. Development and Psychopathology, 19(4), 957–1208. Cicchetti, D. (2010). Resilience under conditions of extreme stress: A multilevel perspective. World Psychiatry, 9, 145–154. Cicchetti, D. (2013). Annual research review: Resilient functioning in maltreated children— past, present, and future perspectives. Journal of Child Psychology and Psychiatry, 54, 402– 422. doi: 10.1111/j.1469–7610.2012.02608.x Cicchetti, D. (Ed.). (2015). Neural plasticity, sensitive periods, and psychopathology. [Special issue]. Development and Psychopathology,27, 319–648.doi: 10.1017/S0954579415000012

Cicchetti, D., & Barnett, D. (1991). Attachment organization in pre-school-aged maltreated children. Development and Psychopathology, 3, 397–411. doi: 10.1017/S0954579400007598 Cicchetti, D., & Blender, J. A. (2006). A multiple-levels-of-analysis perspective on resilience: Implications for the developing brain, neural plasticity, and preventive interventions. Annals of the New York Academy of Sciences, 1094, 248–258. doi: 10.1196/annals.1376.029 Cicchetti, D., & Cannon, T. D. (1999). Neurodevelopmental processes in the ontogenesis and epigenesis of psychopathology. Development and Psychopathology, 11, 375–393. doi: 10.1017/S0954579499002114 Cicchetti, D., & Curtis, W. J. (2006). The developing brain and neural plasticity: Implications for normality, psychopathology, and resilience. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology: Vol. 2. Developmental neuroscience (2nd ed., pp. 1–64). Hoboken, NJ: Wiley. Cicchetti, D., & Curtis, W. J (2007). A multi-level approach to resilience [Special issue]. Development and Psychopathology, 19, 627–955. doi: 10.1017/S0954579407000314 Cicchetti, D., & Dawson, G. (2002). Multiple levels of analysis [Special Issue]. Development and Psychopathology, 14, 417–666. doi: 10.1017/S0954579402003012 Cicchetti, D., & Garmezy, N. (1993). Editorial: Prospects and promises in the study of resilience. Development and Psychopathology, 5, 497–502. doi: 10.1017/S0954579400006118 Cicchetti, D., & Gunnar, M. R. (2008). Integrating biological processes into the design and evaluation of preventive interventions. Development and Psychopathology, 20, 737–743. doi: 10.1017/S0954579408000357 Cicchetti, D., & Hinshaw, S. P. (2002). Prevention and intervention science: Contributions to developmental theory [Special issue]. Development and Psychopathology, 14(4), 667–671. doi: 10.1017.S0954579402004017 Cicchetti, D., & Rogosch, F. A. (1996). Equifinality and multifinality in developmental psychopathology. Development and Psychopathology, 8, 597–600. doi: 10.1017/S0954579400007318 Cicchetti, D., & Rogosch, F. A. (2001a). Diverse patterns of neuroendocrine activity in maltreated children. Development and Psychopathology, 13(3), 677–693. doi: 10.1017/S0954579401003145 Cicchetti, D., & Rogosch, F. A. (2001b). The impact of child maltreatment and psychopathology upon neuroendocrine functioning. Development and Psychopathology, 13, 783–804. Cicchetti, D., & Rogosch, F. A. (2007). Personality, adrenal steroid hormones, and resilience

in maltreated children: A multi-level perspective. Development and Psychopathology, 19(3), 787–809. doi: 10.1017/S0954579407000399 Cicchetti, D., & Rogosch, F. A. (2012). Gene × Environment interaction and resilience: Effects of child maltreatment and serotonin, corticotropin releasing hormone, dopamine, and oxytocin genes. Development and Psychopathology, 24, 411–427. doi: 10.1017/S0954579412000077 Cicchetti, D., Rogosch, F. A., Gunnar, M. R., & Toth, S. L. (2010). The differential impacts of early abuse on internalizing problems and diurnal cortisol activity in school-aged children. Child Development, 81, 252–269. doi: 10.1111/j.1467–8624.2009.01393.x Cicchetti, D., Rogosch, F. A., & Oshri, A. (2011). Interactive effects of corticotropin releasing hormone receptor 1, serotonin transporter linked polymorphic region, and child maltreatment on diurnal cortisol regulation and internalizing symptomatology. Development and Psychopathology, 23, 1125–1138. doi: 10.1017/S0954579411000599 PMCID 22018085 Cicchetti, D., Rogosch, F. A., & Toth, S. L. (2006). Fostering secure attachment in infants in maltreating families through preventive interventions. Development and Psychopathology, 18(3), 623–650. doi: 10.1017/S0954579406060329 Cicchetti, D., Rogosch, F. A., Toth, S. L., & Sturge-Apple, M. L. (2011). Normalizing the development of cortisol regulation in maltreated infants through preventive interventions. Development and Psychopathology, 23, 789–800. doi: 10.1017/S0954579411000307 Cicchetti, D., & Roisman, G. I. (Eds.) (2011). The origins and organization of adaptation and maladaptation: Minnesota symposia on child psychology. Hoboken, NJ: Wiley. doi: 10.1002/9781118036600 Cicchetti, D., & Toth, S. L. (1995). Developmental psychopathology and disorders of affect. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Risk, disorder, and adaptation (Vol. 2, pp. 369–420). New York, NY: Wiley. Cicchetti, D., & Toth, S. L. (2006). Translational research and developmental psychopathology [Special Issue]. Development and Psychopathology, 18(3), 619–933. doi: 10.1017/S0954579406060317 Cicchetti, D., & Toth, S. L. (2009). The past achievements and future promises of developmental psychopathology: The coming of age of a discipline. Journal of Child Psychology and Psychiatry, 50, 16–25. doi: 10.1111/j.1469–7610.2008.01979.x Cicchetti, D., & Toth, S. L. (2015). Child maltreatment. In In R. M. Lerner (Ed.), M. Lamb & C. García Coll (Vol. Eds.), Handbook of child psychology and developmental science, Vol. 3: Socioemotional process (7th ed.; pp. 513–563). Hoboken, NJ: Wiley. Cicchetti, D., Toth, S. L., & Handley, E. D. (2015). Genetic moderation of interpersonal psychotherapy efficacy for low-income mothers with major depressive disorder: Implications for differential susceptibility. Development and Psychopathology, 27, 19–35. doi:

10.1017/S0954579414001278 Cicchetti, D., Toth, S. L., & Rogosch, F. A. (1999). The efficacy of toddler-parent psychotherapy to increase attachment security in offspring of depressed mothers. Attachment and Human Development, 1, 34–66. doi: 10.1080/14616739900134021 Cicchetti, D., & Tucker, D. (1994). Development and self-regulatory structures of the mind. Development and Psychopathology, 6, 533–549. doi: 10.1080/14616739900134021 Cicchetti, D., & Walker, E. F. (Eds.). (2003). Neurodevelopmental mechanisms in psychopathology. New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511546365 Coddington, R. D. (1972a). The significance of life events as etiologic factors in the diseases of children: I. A survey of professional workers. Journal of Psychosomatic Research, 16, 7– 18. doi: 10.1016/0022–3999(72)90018–9 Coddington, R. D. (1972b). The significance of life events as etiologic factors in the diseases of children: II. A study of a normal population. Journal of Psychosomatic Research, 16, 205– 213. doi: 10.1016/0022–3999(72)90045–1 Cohen, J. (2013). Creating a positive school climate: A foundation for resilience. In S. Goldstein & R. B. Brooks (Eds.), Handbook of resilience in children (2nd ed., pp. 411–425). New York, NY: Springer. doi: 10.1007/978–1–4614–3661–4_24 Cole, D. A., & Maxwell, S. E. (2003). Testing mediational models with longitudinal data: Questions and tips in the use of structural equation modeling. Journal of Abnormal Psychology, 112(4), 558–577. doi: 10.1037/0021–843X.112.4.558 Collins, F. S., Morgan, M., & Patrinos, A. (2003). The Human Genome Project: Lessons from large-scale biology. Science, 300, 286–290. doi: 10.1126/science.1084564 Collishaw, S., Pickles, A., Messer, J., Rutter, M., Shearer, C., & Maughan, B. (2007). Resilience to adult psychopathology following childhood maltreatment: Evidence from a community sample. Child Abuse and Neglect, 31, 211–229. doi: 10.1016/j.chiabu.2007.02.004 Colville, G., & Cream, P. (2009). Post-traumatic growth in parents after a child's admission to intensive care: Maybe Nietzsche was right? Intensive care medicine, 35(5), 919–923. doi: 10.1007/s00134–009–1444–1 Compas, B. E. (1987). Stress and life events during childhood and adolescence. Clinical Psychology Review, 7(3), 275–302. doi: 10.1016/0272–7358(87)90037–7 Conger, K. J., Rueter, M. A., & Conger, R. D. (2000). The role of economic pressure in the lives of parents and their adolescents: The family stress model. In L. J. Crockett & R. K. Silbereisen (Eds.), Negotiating adolescence in times of social change (pp. 201–223). New

York, NY: Cambridge University Press. Conger, R. D., & Elder, G. H., Jr. (1994). Families in troubled times: Adapting to change in rural America. Hawthorne, NY: de Gruyter. Cowan, W. M., Kopnisky, K. L., & Hyman, S. E. (2002). The Human Genome Project and its impact on psychiatry. Annual Review of Neuroscience, 25, 1–50. doi: 10.1146/annurev.neuro.25.112701.142853 Cowell, R. A. Cicchetti, D, Rogosch, F. A. & Toth, S. L. (2015). Childhood maltreatment and its effect on neurocognitive functioning: Chronicity and timing matter. Development and Psychopathology, 27, 521–533. doi: 10.1017/S0954579415000139 Cozza, S. J., & Lerner, R. M. (Eds.). (2013). Military children and families: Introducing the issue [Special issue]. Future of Children, 23, 3–11. Crawford, D. C., & Nickerson, D. A. (2005). Definition and clinical importance of haplotypes. Annual Review of Medicine, 56, 303–320. doi: 10.1146/annurev.med.56.082103.104540 Crawford, E., Wright, M. O., & Masten, A. S. (2006). Resilience and spirituality in youth. In E. C. Roehlkepartain, P. E. King, L. Wagener, & P. L. Benson (Eds.), The handbook of spiritual development in childhood and adolescence (pp. 355–370). Thousand Oaks, CA: Sage. doi: 10.4135/9781412976657.n25 Cunha, F., & Heckman, J. J. (2011). The economics and psychology of inequality and human development. Journal of the European Economic Association, 7, 320–364. doi: 10.1162/JEEA.2009.7.2–3.320 Curtis, W. J., & Cicchetti, D. (2003). Moving research on resilience into the 21st century: Theoretical and methodological considerations in examining the biological contributors to resilience. Development and Psychopathology, 15, 773–810. doi: 10.1017/S0954579403000373 Curtis W. J., & Cicchetti, D. (2007). Emotion and resilience: A multi-level investigation of hemispheric electroencephalogram asymmetry and emotion regulation in maltreated and nonmaltreated children. Development and Psychopathology, 19(3), 811–840. doi: 10.1017/S0954579407000405 Curtis, W. J., & Nelson, C. A. (2003). Resilience and vulnerability: Adaptation in the context of childhood adversities. In S. Luthar (Ed.), Toward building a better brain: Neurobehavioral outcomes, mechanisms, and processes of environmental enrichment (pp. 463–488). New York, NY: Cambridge University Press. Cyr, C., Euser, E. M., Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2010). Attachment security and disorganization in maltreating and high-risk families: A series of metaanalyses. Development and Psychopathology, 22, 87–108. doi: 10.1017/S0954579409990289

D'Argembeau, A., Cassol, H., Phillips, C., Balteau, E., Salmon, E., & Van der Linden, M. (2014). Brains creating stories of selves: The neural basis of autobiographical reasoning. Social cognitive and affective neuroscience, 9, 646–652. doi: 10.1093/scan/nst028 Dackis, M. N., Rogosch, F. A., Oshri, A., & Cicchetti, D. (2012). The role of limbic system irritability in linking history of childhood maltreatment and psychiatric outcomes in low income, high-risk women: Moderation by FKBP5. Development and Psychopathology, 24(4), 1237–1252. doi: 10.1017/S0954579412000673 Danese, A., Moffitt, T. E., Pariante, C. M., Ambler, A., Poulton, R., & Caspi, A. (2008). Elevated inflammation levels in depressed adults with a history of childhood maltreatment. Archives of General Psychiatry, 65, 409–416. doi: 10.1001/archpsyc.65.4.409 Davidson, R. J. (2000). Affective style, psychopathology, and resilience: Brain mechanisms and plasticity. American Psychologist, 55, 1196–1214. doi: 10.1037/0003–066X.55.11.1196 Davis-Kean, P. E., & Jager, J. (2014). Trajectories of achievement within race/ethnicity: “Catching up” I achievement across time. Journal of Educational Research, 107, 197–208. doi: 10.1080/00220671.2013.807493 DeBellis, M. D. (2001). Developmental traumatology: The psychobiological development of maltreated children and its implications for research, treatment, and policy. Development and Psychopathology, 13, 539–564. doi: 10.1017/S0954579401003078 DeBellis, M. D. (2005). The psychobiology of neglect. Child Maltreatment, 10, 150–172. doi: 10.1177/1077559505275116 Del Giudice, M., Ellis, B. J., & Shirtcliff (2011). The adaptive calibration model of stress responsivity. Neuroscience and Biobehavioral Reviews, 35, 1562–1592. doi: 10.1016/j.neubiorev.2010.11.007 DeYoung, C., Cicchetti, D., & Rogosch, F. A. (2011). Moderation of the association between childhood maltreatment and neuroticism by the corticotropin-releasing hormone receptor 1 gene. Journal of Child Psychology and Psychiatry, 52, 898–906. doi: 10.1111/j.1469– 7610.2011.02404.x DeYoung, C. G., Cicchetti, D., Rogosch, F. A., Gray, J. R., Eastman, M., & Grigorenko, E. L. (2011). Sources of cognitive exploration: Genetic variation in the prefrontal dopamine system predicts openness/intellect. Journal of Research in Personality, 45, 364–371. doi: 10.1016/j.jrp.2011.04.002 Diamond, A., & Lee, K. (2011). Interventions shown to aid executive function development in children 4 to 12 years old. Science, 333, 959–964. doi: 10.1126/science.1204529 Dishion, T. J., & Tipsord, J. M. (2011). Peer contagion in child and adolescent social and emotional development. Annual Review of Psychology, 62, 189–214. doi: 10.1146/annurev.psych.093008.100412

Dodge, K. A., Greenberg, M. T., Malone, P. S., & the Conduct Problems Prevention Research Group (2008). Testing an idealized dynamic cascade model of the development of serious violence in adolescence. Child Development, 79, 1907–1927. doi: 10.1111/j.1467– 8624.2008.01233.x Doll, B. (2013). Enhancing resilience in classrooms. In S. Goldstein & R. B. Brooks (Eds.), Handbook of resilience in children (pp. 399–410). New York, NY: Springer. doi: 10.1007/978–1–4614–3661–4_23 Doll, B., LeClair, C., & Kurien, S. (2009). Effective classrooms: Classroom learning environments that foster school success. In T. B. Gutkin & C. R. Reynolds (Eds.), The handbook of school psychology (4th ed., pp. 791–807). Hoboken, NJ: Wiley. Doom, J. R., & Gunnar, M. R. (2013). Stress physiology and developmental psychopathology: Past, present, and future. Development and Psychopathology, 25(4, pt. 2), 1359–1373. doi: 10.1017/S0954579413000667 Dozier, M., Peloso, E., Lewis, E., Laurenceau, J.-P., & Levine, S. (2008). Effects of an attachment-based intervention on the cortisol production of infants and toddlers in foster care. Developmental Psychopathology, 20, 845–859. doi: 10.1017/S0954579408000400 Dubowitz, H., & Poole, G. (2012). Child neglect: An overview. Encyclopedia on early childhood development. Retrieved from www.child-encyclopedia.com/ pages/PDF/DubowitzPoolANGxp1.pdf Duckworth, A. L. (2011). The significance of self-control. Proceedings of the National Academy of Sciences, 108, 2639–2640. doi: 10.1073/pnas.1019725108 Duncan, G. J. (2012). Give us this day our daily breadth. Child Development, 83, 6–15. doi: 10.1111/j.1467–8624.2011.01679.x Duncan, L. E., & Keller, M. C. (2011). A critical review of the first 10 years of candidate gene-by-environment interaction research in psychiatry. American Journal of Psychiatry, 168, 1041–1049. doi: 10.1176/appi.ajp.2011.11020191 Duncan, L. E., Pollastri, A. R., & Smoller, J. W. (2014). Mind the gap: Why geneticists and psychological scientists have discrepant views about gene-environment interaction (G×E) research. American Psychologist, 69, 249–268. doi: 10.1037/a0036320. Durlak, J. A., Weissberg, R. P., Dymnicki, A. B., Schellinger, K. B., & Taylor, R. D. (2011). The impact of enhancing students' social and emotional learning: A meta-analysis of schoolbased universal interventions. Child Development, 82, 405–432. doi: 10.1111/j.1467– 8624.2010.01564.x Durlak, J. A., Weissberg, R. P., & Pachan, M. (2010). A meta-analysis of after-school programs that seek to promote personal and social skills in children and adolescents. American Journal of Community Psychology, 45, 294–309. doi: 10.1007/s10464–010–9300–

6 Eccles, J. S., & Roeser, R. W. (2012). School influences on human development. In L. C. Mayes & M. Lewis (Eds.), The Cambridge handbook of environment in human development (pp. 259–283). New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139016827.017 Egeland, B., Carlson, E., & Sroufe, L. A. (1993). Resilience as process. Development and Psychopathology, 5, 517–528. doi: 10.1017/S0954579400006131 Eggerman, M., & Panter-Brick, C. (2010). Suffering, hope, and entrapment: Resilience and cultural values in Afghanistan. Social Science and Medicine, 71, 71–83. doi: 10.1016/j.socscimed.2010.03.023 Eisenberg, L. (1995). The social construction of the human brain. American Journal of Psychiatry, 152, 1563–1575. doi: 10.1176/ajp.152.11.1563 Elder, G. H., Jr., & Conger, R. D. (2000). Children of the land: Adversity and success in rural America. Chicago, IL: University of Chicago Press. Ellis, B. J., Boyce, W. T., Belsky, J., Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2011). Differential susceptibility to the environment: An evolutionary–neurodevelopmental theory. Development and Psychopathology, 23, 7–28. doi: 10.1017/S0954579410000611 Ertl, V., Pfeiffer, A., Schauer, E., Elbert, T., & Neuner, F. (2011). Community-implemented trauma therapy for former child soldiers in northern Uganda: A randomized controlled trial. Journal of the American Medical Association, 306, 503–512. doi: 10.1001/jama.2011.1060 Evans, A. B., Banerjee, M., Meyer, R., Aldana, A., Foust, M., & Rowley, S. (2012). Racial socialization as a mechanism for positive development among African American youth. Child Development Perspectives, 6, 251–257. doi: 10.1111/j.1750–8606.2011.00226.x Evans, G. W., Li, D., & Sepanski Whipple, S. (2013). Cumulative risk and child development. Psychological Bulletin, 139, 1342–1396. doi: 10.1037/a0031808 Evans, G. W., & Wachs, T. D. (2009). Chaos and its influence on children's development. Washington DC: American Psychological Association. Feder, A., Charney, D., & Collins, K. (2011). Neurobiology of resilience. In S. M. Southwick, B. T. Litz, D. Charney, & M. J. Friedman (Eds.), Resilience and mental health: Challenges across the lifespan (pp. 1–29). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511994791.003 Feder, A., Nestler, E. J., & Charney, D. S. (2009). Psychobiology and molecular genetics of resilience. Nature Reviews Neuroscience, 10, 446–457. doi: 10.1038/nrn2649 Felitti, V. J., Anda, R. F., Nordenberg, D., Williamson, D. F., Spitz, A. M., Edwards, V.,… Marks, J. S. (1998). Relationships of childhood abuse and household dysfunction to many of

the leading causes of death in adults: The Adverse Childhood Experiences (ACE) study. American Journal of Preventive Medicine, 14, 245–258. doi: 10.1016/S0749– 3797(98)00017–8 Fiese, B. H. (2006). Family routines and rituals. New Haven, CT: Yale University Press. Figueiredo, C. A., Amorim, L. D., Alcantara-Neves, N. M., Matos, S. M. A., Cooper, P. J., Rodriques, L. C., & Barreto, M. L. (2013). Environmental conditions, immunologic phenotypes, atopy, and asthma: New evidence of how the hygiene hypothesis operates in Latin America. Journal of Allergy and Clinical Immunology, 131, 1064–1068. doi: 10.1016/j.jaci.2013.01.016 Fishbein, D. (2000). The importance of neurobiological research to the prevention of psychopathology. Prevention Science, 1, 89–106. doi: 10.1023/A:1010090114858 Fisher, C. B., Busch-Rossnagel, N. A., Jopp, D. S., & Brown, J. L. (2012). Applied developmental science, social justice, and socio-political well-being. Applied Developmental Science, 16(1), 54–64. doi: 10.1080/10888691.2012.642786 Fisher, P. A., Gunnar, M. R., Dozier, M., Bruce, J., & Pears, K. C. (2006). Effects of therapeutic interventions for foster children on behavioral problems, caregiver attachment, and stress regulatory neural systems. Annals of the New York Academy of Sciences, 1094, 215– 225. doi: 10.1196/annals.1376.023 Fisher, P. A., Stoolmiller, M., Gunnar, M. R., & Burraston, B. O. (2007). Effects of a therapeutic intervention for foster preschoolers on daytime cortisol activity. Psychoneuroendocrinology, 32, 892–905. doi: 10.1016/j.psyneuen.2007.06.008 Fisher, P. A., Van Ryzin, M. J., & Gunnar, M. R. (2011). Mitigating HPA axis dysregulation associated with placement changes in foster care. Psychoneuroendocrinology, 36, 531–539. doi: 10.1016/j.psyneuen.2010.08.007 Flynn, R. J., Dudding, P. M., & Barber, J. G. (Eds.). (2006). Promoting resilience in child welfare. Ottawa, Canada: University of Ottawa Press. Folke, C., Carpenter, S. R., Walker, B., Scheffer, M., Chapin, T., & Rockström, J. (2010). Resilience thinking: integrating resilience, adaptability and transformability. Ecology and Society, 15, 20 [Online]. Retrieved from http://www.ecologyandsociety.org/vol15/iss4/art20/ Fox, N. A., Almas, A. N., Degnan, K. A., Nelson, C. A., & Zeanah, C. H. (2011). The effects of severe psychosocial deprivation and foster care intervention on cognitive development at 8 years of age: Findings from the Bucharest Early Intervention Project. Journal of Child Psychology and Psychiatry, 52, 919–928. doi: 10.1111/j.1469–7610.2010.02355.x Fox, N. A., Nelson, C. A., & Zeanah, C. H. (2014). The problem of institutionalization of young children and its consequences for efforts to build peaceful societies.In J. F. Leckman, C. Panter-Brick, & R. Salah (Eds.), Pathways to peace: The transformative power of families

and child development (pp. 145–159). Cambridge MA: MIT Press. Frankl, V. E. (2006). Man's search for meaning. Boston: Beacon Press. (Original work published 1959). Fuligni, A. J., & Telzer, E. H. (2012). The contributions of youth to immigrant families. In A. S. Masten, K. Liebkind, & D. J. Hernandez (Eds.), Realizing the potential of immigrant youth (pp. 181–202). New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139094696.010 Furr, J. M., Comer, J. S., Edmunds, J. M., & Kendall, P. C. (2010). Disasters and youth: A meta-analytic examination of posttraumatic stress. Journal of Consulting and Clinical Psychology, 78, 765–780. doi: 10.1037/a0021482 Garmezy, N. (1974). The study of competence in children at risk for severe psychopathology. In E. J. Anthony & C. Koupernik (Eds.), The child in his family: Children at psychiatric risk (Vol. 3, pp. 77–98). New York, NY: Wiley. Garmezy, N. (1981). Children under stress: Perspectives on antecedents and correlates of vulnerability and resistance to psychopathology. In A. I. Rabin, J. Aronoff, A. M. Barclay, & R. A. Zucker (Eds.), Further explorations in personality (pp. 196–269). New York, NY: Wiley. Garmezy, N. (1982a). Foreword. In E. E. Werner & R. S. Smith (Eds.), Vulnerable but invincible: A longitudinal study of resilient children and youth (pp. xiii–xix). New York, NY: McGraw-Hill. Garmezy, N. (1982b). The case for the single case in research. In A. E. Kazdin & A. H. Tuma (Eds.), New directions for methodology of social and behavioral sciences: Single-case research designs (pp. 5–17). San Francisco: Jossey-Bass. Garmezy, N. (1985). Stress-resistant children: The search for protective factors. In J. E. Stevenson (Ed.), Recent research in developmental psychopathology: Journal of child psychology and psychiatry book supplement, No. 4 (pp. 213–233). Oxford, UK: Pergamon Press. Garmezy, N., Masten, A. S., & Tellegen, A. (1984). The study of stress and competence in children: A building block for developmental psychopathology. Child Development, 55, 97– 111. doi: 10.2307/1129837 Garmezy, N., & Rutter, M. (Eds.). (1983). Stress, coping, and development in children. New York, NY: McGraw-Hill. Gee, G. D., Gabard-Durnam, L. J., Flannery, J., Goff, B., Humphreys, K. L., Telzer, E. H.,… Tottenham, N. (2013). Early developmental emergence of human amygdala-prefrontal connectivity after maternal deprivation. Proceedings of the National Academy of Sciences of the United States of America, 110(39), 15638–15643. doi: 10.1073/pnas.1307893110

Gest, S. D., & Davidson, A. J. (2011). A developmental perspective on risk, resilience, and prevention. In M. Underwood & L. Rosen (Eds.), Social development: Relationships in infancy, childhood, and adolescence (pp. 427–454). New York, NY: Guilford Press. Gest, S. D., Reed, M.-G. J., & Masten, A. S. (1999). Measuring developmental changes in exposure to adversity: A life chart and rating scale approach. Development and Psychopathology, 11, 171–192. doi: 10.1017/S095457949900200X Glueck, S., & Glueck, E. (1950). Unraveling juvenile delinquency. Cambridge, MA: Harvard University Press. Goldenberg, H., & Goldenberg, I. (2013). Family therapy: An overview (8th ed.). Belmont, CA: Brooks/Cole. Goldstein, S., & R. B. Brooks (2013). Handbook of resilience in children (2nd ed.). New York, NY: Springer. doi: 10.1007/978–1–4614–3661–4 Goodyer, I., Kolvin, I., & Gatzanis, S. (1985). Recent undesirable life events and psychiatric disorder in childhood and adolescence. British Journal of Psychiatry, 147, 517–523. doi: 10.1192/bjp.147.5.517 Gottesman, I. I. (1974). Developmental genetics and ontogenetic psychology: Overdue détente and propositions from a matchmaker. In A. Pick (Ed.), The Minnesota symposia on child psychology, (Vol. 8, pp. 55–80). Minneapolis, MN: University of Minnesota Press. Gottesman, I. I., & Gould, T. D. (2003). The endophenotype concept in psychiatry: Etymology and strategic intentions. American Journal of Psychiatry, 160, 636–645. doi: 10.1176/appi.ajp.160.4.636 Gottesman, I. I., & Hanson, D. R. (2005). Human development: Biological and genetic processes. Annual Review of Psychology, 56, 263–286. doi: 10.1146/annurev.psych.56.091103.070208 Gottesman, I., & Shields, J. (1972). Schizophrenia and genetics: A twin study vantage point. New York, NY: Academic Press. Gottesman, I. I., & Shields, J. (1982) Schizophrenia: The epigenetic puzzle. Cambridge, UK: Cambridge University Press. Gottlieb, G. (2007). Probabilistic epigenesis. Developmental Science, 10(1), 1–11. doi: 10.1111/j.1467–7687.2007.00556.x Greenough, W., Black, J., & Wallace, C. (1987). Experience and brain development. Child Development, 58, 539–559. doi: 10.1111/1467–8624.ep7264422 Grigorenko, E. L., & Cicchetti, D. (2012a) Genomic sciences for developmentalists: The current state of affairs. Development and Psychopathology, 24, 1157–1164. doi: 10.1017/S09545794120006128

Grigorenko, E. L., & Cicchetti, D. (Eds.) (2012b). The contribution of genetic/genomic sciences to developmental psychopathology [Special issue]. Development and Psychopathology, 24, 1157–1451. doi: 10.1017/S0954579412000612 Guerra, S., & Martinez, F. D. (2008). Asthma genetics: From linear to multifactorial approaches. Annual Review of Medicine, 59, 327–341. doi: 10.1146/annurev.med.59.060406.213232 Gunderson, L. (2010). Ecological and human community resilience in response to natural disasters. Ecology and Society, 15. Retrieved from www. ecologyandsociety.org/vol15/iss2/art18/ Gunnar, M., & Fisher, P. A., (2006). The Early Experience, Stress, and Prevention Network. Bringing basic research on early experience and stress neurobiology to bear on preventive interventions for neglected and maltreated children. Development and Psychopathology, 18(3), 651–678. doi: 10.1017/S0954579406060330 Gunnar, M. R., & Herrare, A. M. (2013). The neurobiology of stress and development. In P. D. Zelazo (Ed.), Oxford handbook of developmental psychology (Vol. 2; pp. 45–80).New York, NY: Oxford University Press. Gunnar, M. R., & Vazquez, D. M. (2001). Low cortisol and a flattening of expected daytime rhythm: Potential indices of risk in human development. Development and Psychopathology, 13, 515–538. doi: 10.1017/S0954579401003066 Gunnar, M. R., Wenner, J. A., Thomas, K. M., Glatt, C. E., McKenna, M. C., & Clark, A. G. (2012). The brain-derived neurotrophic factor Val66Met polymorphism moderates early deprivation effects on attention problems. Development and Psychopathology, 24(4), 1215– 1223. doi: 10.1017/S095457941200065X Gutman, L. M., Sameroff, A. J., and Eccles, J. S. (2002). The academic achievement of African American students during early adolescence: An examination of multiple risk, promotive, and protective factors. American Journal of Community Psychology, 30, 367–399. doi: 10.1023/A:1015389103911 Hanson, D., & Gottesman, I. I. (2007). Choreographing genetic, epigenetic, and stochastic steps in the dances of developmental psychopathology. In A. S. Masten (Ed.), Multilevel dynamics in developmental psychopathology: Pathways to the future: The Minnesota symposia on child psychology, Vol. 34 (pp. 27–44). Mahwah, NJ: Erlbaum. Harkness, S., & Super, C. M. (2012). The cultural organization of children's environments. In L. C. Mayes & M. Lewis (Eds.), The Cambridge handbook of environment in human development (pp. 498–516). New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139016827.028 Hart, S. (2012). Siblings and peers in the adult-child-child triadic context. In L. C. Mayes &

M. Lewis (Eds.), The Cambridge handbook of environment in human development (pp. 284– 299).New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139016827.018 Hart, H., & Rubia, K. (2012). Neuroimaging of child abuse: a critical review. Frontiers in Human Neuroscience, 6(52). doi: 10.3389/fnhum.2012.00052. Hartup, W. W. (1996). The company they keep: Friendships and their developmental significance. Child Development, 67, 1–13. doi: 10.1111/j.1467–8624.1996.tb01714.x Hartup, W. W., & Stevens, N. (1997). Friendships and adaptation in the life course. Psychological Bulletin, 121, 355–370. doi: 10.1037/0033–2909.121.3.355 Hauser, S. T., Allen, J. P., & Golden, E. (2006). Out of the woods: Tales of resilient teens. Cambridge, MA: Harvard University Press. Hawkins, J. D., Kosterman, R., Catalano, R. F., Hill, K. G., & Abbott, R. D. (2005). Promoting positive adult functioning through social development intervention in childhood: Long-term effects from the Seattle Social Development Project. Archives of Pediatrics and Adolescent Medicine, 159, 25–31. doi: 10.1001/archpedi.159.1.25 Hawley, D. R., & DeHaan, L. (1996). Toward a definition of family resilience: Integrating life-span and family perspectives. Family Process, 35, 283–298. doi: 10.1111/j.1545– 5300.1996.00283.x Heckman, J. J. (2006). Skill formation and the economics of investing in disadvantaged children. Science, 312, 1900–1902. doi: 10.1126/science.1128898 Heim, C., Newport, J. D., Mletzko, T., Miller, A. H., & Nemeroff, C. B. (2008). The link between childhood trauma and depression: Insights from HPA axis studies in humans. Psychoneuroendocrinology, 33, 693–710. doi: 10.1016/j.psyneuen.2008.03.008. Herbers, J. E., Cutuli, J. J., Supkoff, L. M., Narayan, A. J., & Masten, A. S. (2014). Parenting and co-regulation: Adaptive systems for competence in children experiencing homelessness. American Journal of Orthopsychiatry, 84, 420–430. doi: 10.1037/h0099843 Hetherington, E. M. (1979). Divorce: a child's perspective. American Psychologist, 34(10), 851–858. doi: 10.1037/0003–066X.34.10.851 Hinshaw, S. P. (1992). Externalizing behavior problems and academic underachievement in childhood and adolescence: causal relationships and underlying mechanisms. Psychological Bulletin, 111, 127–155. doi: 10.1037/0033–2909.111.1.127 Hinshaw, S. P., & Anderson, C. A. (1996). Conduct and oppositional defiant disorders. In E. J. Mash & R. A. Barkley (Eds.), Child psychopathology (pp. 113–149). New York, NY: Guilford Press. Hinshaw, S. P., & Cicchetti, D. (2000). Stigma and mental disorder: Conceptions of illness, public attitudes, personal disclosure, and social policy. Development and Psychopathology,

12, 555–598. doi: 10.1017/S0954579400004028 Holmes, T. H., & Rahe, R. H. (1967). The social readjustment rating scale. Journal of Psychosomatic Research, 11, 213–218. doi: 10.1016/0022–3999(67)90010–4 Hostinar, C. E., Cicchetti, D., & Rogosch, F. A. (2014). Oxytocin receptor gene (OXTR) polymorphism, perceived social support, and psychological symptoms in maltreated adolescents. Development and Psychopathology, 26, 465–477. doi: 10.1017/S0954579414000066. Hostinar, C. E., Sullivan, R. M., & Gunnar, M. R. (2014). Psychobiological mechanisms underlying the social buffering of the hypothalamic–pituitary–adrenocortical axis: A review of animal models and human studies across development. Psychological Bulletin, 140(1), 256– 282. doi: 10.1037/a0032671 Howe, G. W., Beach, S. R., & Brody, G. H. (2010). Mitochondrial methods for translating gene-environment dynamics into preventive interventions. Prevention Science, 11(4), 343– 354. doi: 10.1007/s11121–010–0177–2 Howe, G. W., Reiss, D., & Yuh, J. (2002). Can prevention trials test theories of etiology? Development and Psychopathology, 14, 673–694. doi: 10.1017/S0954579402004029 Hughes, D., Rodriguez, J., Smith, E. P., Johnson, D. J., Stevenson, H. C., & Spicer, P. (2006). Parents' ethnic-racial socialization practices: A review of research and directions for future study. Developmental Psychology, 42, 747–770. doi: 10.1037/0012–1649.42.5.747 Insel, T. R., & Quiron, R. (2005). Psychiatry as a clinical neuroscience discipline. Journal of the American Medical Association, 294, 2221–2224. doi: 10.1001/jama.294.17.2221 Institute of Medicine (2014). Preparedness, response, and recovery considerations for children and families: Workshop summary. Washington, DC: National Academies Press. Jordan, L. P., & Graham, E. (2012). Resilience and well-being among children of migrant parents in South-East Asia. Child Development, 83, 1672–1688. doi: 10.1111/j.1467– 8624.2012.01810.x Jung, T., & Wickrama, K. A. S. (2008). An introduction to latent class growth analysis and growth mixture modeling. Social and Personality Psychology Compass, 2, 302–317. doi: 10.1111/j.1751–9004.2007.00054.x Juster, R. P., Bizik, G., Picard, M., Arsenault-Lapierre, G., Sindi, S., Trepanier, L.,…Lupien, S. J. (2011). A transdisciplinary perspective of chronic stress in relation to psychopathology throughout lifespan development. Development and Psychopathology, 23(3), 725–776. doi: 10.1017/S0954579411000289 Juster, R. P., McEwen, B. S., & Lupien, S. J. (2009). Allostatic load biomarkers of chronic stress and impact on health and cognition. Neuroscience and Biobehavioral Reviews, 35, 2–

16. doi: 10.1016/j.neubiorev.2009.10.002 Kagan, J. (2005). Human morality and temperament. Nebraska Symposium on Motivation, 51, 1–32. Kandel, E. R. (1998). A new intellectual framework for psychiatry. American Journal of Psychiatry, 155, 457–469. doi: 10.1016/S0014–3855(02)00104–4 Karatsoreos, I. N., & McEwen, B. S. (2011). Psychobiological allostasis: Resistance, resilience and vulnerability. Trends in Cognitive Neuroscience, 15, 576–584. doi: 10.1016/j.tics.2011.10.005 Karatsoreos, I. N., & McEwen, B. S. (2013). Annual research review: The neurobiology and physiology of resilience and adaptation across the life course. Journal of Child Psychology and Psychiatry, 54, 337–347. doi: 10.1111/jcpp.12054 Karg, K., Burmeister, M., Shedden, K., & Sen, S. (2011). The serotonin transporter promoter variant (5-HTTLPR), stress, and depression meta-analysis revisited: Evidence of genetic moderation. Archives of General Psychiatry, 68(5), 444–454. doi: 10.1001/archgenpsychiatry.2010.189 Kaufman, J., Yang, B., Douglas-Palumberi, H., Grasso, D., Lipschitz, D, Houshyar, S.,… Gelertner, J. (2006). Brain-derived neurotrophic factor-5-HTTLPR gene interactions and environmental modifiers of depression in children. Biological Psychiatry, 59, 673–680. doi: 10.1016/j.biopsych.2005.10.026 Kaufman, J., Yang, B., Douglas-Palumberi, H., Houshyar, S., Lipschitz, D., Krystal, J., & Gelernter, J. (2004). Social supports and serotonin transporter gene moderate depression in maltreated children. Proceedings of the National Academy of Sciences of the United States of America, 101(49), 17316–17321. doi: 10.1073/pnas.0404376101 Keller, M. C. (2014). Gene × environment interaction studies have not properly controlled for potential confounders: The problem and the (simple) solution. Biological Psychiatry, 75, 18– 24. doi: 10.1016/j.biopsych.2013.09.006 Kelly, J. B., & Emery, R. E. (2003). Children's adjustment following divorce: Risk and resilience perspectives. Family Relations, 52(4), 352–362. doi: 10.1111/j.1741– 3729.2003.00352.x Kim-Cohen, J., & Gold, A. L (2009). Measured gene-environment interactions and mechanisms promoting resilient development. Current Directions in Psychological Science, 18, 138–142. doi: 10.1111/j.1467–8721.2009.01624.x Kim-Cohen, J., & Turkewitz, R. (2012). Resilience and measured gene–environment interactions. Development and Psychopathology, 24, 1297–1306. doi: 10.1017/S0954579412000715

King, L. A., Hicks, J. A., Krull, J. L., & Del Gaiso, A. K. (2006). Positive affect and the experience of meaning in life. Journal of Personality and Social Psychology, 90, 179–196. doi: 10.1037/0022–3514.90.1.179 Kirschman, K. J. B., Johnson, R. J., Bender, J. A., & Roberts, M. C. (2009). Positive psychology for children and adolescents: Development, prevention and promotion. In C. R. Snyder & S. Lopez (Eds.), Oxford Handbook of positive psychology (pp. 133–148). New York, NY: Oxford University Press. Klasen, F., Oettingen, G., Daniels, J., Post, M., Hoyer, C., & Adam, H. (2010). Posttraumatic resilience in former Ugandan child soldiers. Child Development, 81, 1096–1113. doi: 10.1111/j.1467–8624.2010.01456.x Kobasa, S. C. (1979). Stressful life events, personality, and health: An inquiry into hardiness. Journal of Personality and Social Psychology, 37, 1–11. doi: 10.1037/0022–3514.37.1.1 Kohlberg, L., LaCrosse, J., & Ricks, D. (1972). The predictability of adult mental health from childhood behavior. In B. B. Wolman (Ed.), Manual of child psychopathology (pp. 1217– 1284). New York, NY: McGraw-Hill. Kolb, B., & Gibb, R. (2001). Early brain injury, plasticity, and behavior. In C. A. Nelson & M. Luciana (Eds.), Handbook of developmental cognitive neuroscience (pp. 175–190). Cambridge, MA: The MIT Press. Kolb, B., & Whishaw, I. Q. (1998). Brain plasticity and behavior. Annual Review of Psychology, 49, 43–64. doi: 10.1146/annurev.psych.49.1.43 Kopp, C. B. (1983). Risk factors in development. In M. M. Haith & J. J. Campos (Eds.), Handbook of child psychology: Vol. 2. Infancy and developmental psychobiology (4th ed., pp. 1081–1188). New York, NY: Wiley. Ladd, G. W. (2005). Children's peer relations and social competence: A century of progress. New Haven, CT: Yale University Press. La Greca, A. M., Lai, B. S., Joormann, J., Auslander, B. B., & Short, M. A. (2013). Children's risk and resilience following a natural disaster: Genetic vulnerability, posttraumatic stress, and depression. Journal of Affective Disorders, 151, 860–867. doi: 10.1016/j.jad.2013.07.024 Latif, F., Yeatermeyer, J., Horne, Z. D., & Beriwal, S. (in press). Psychological impact of nuclear disasters in children and adolescents. Child and Adolescent Psychiatric Clinics of North America. doi: 10.106/j.chc.2015.06.009 Laub, J. H., & Sampson, R. J. (2003). Shared beginnings, divergent lives: Delinquent boys to age 70. Cambridge, MA: Harvard University Press. Lauer, P. A., Akiba, M., Wilkerson, S. B., Apthorp, H. S., Snow, D., & Martin-Glenn, M. L. (2006). Out-of-school-time programs: A meta-analysis of effects for at-risk students. Review of

Educational Research, 76, 275–313. doi: 10.3102/00346543076002275 Layne, C. M., Olsen, J. A., Baker, A., Legerski, J.-P., Isakson, B., Pašalić, A…, Pynoos, R. S. (2010). Unpacking trauma exposure risk factors and differential pathways of influence: Predicting postwar mental distress in Bosnian adolescents. Child Development, 81, 1053– 1076. doi: 10.1111/j.1467–8624.2010.01454.x Leckman, J. F., Panter-Brick, C., & Salah, R. (2014). Pathways to peace: The transformative power of children and families. Cambridge, MA: MIT Press. Lenroot, R. K. & Giedd, J. N. (2008). The changing impact of genes and environment on brain development during childhood and adolescence: Initial findings from a neuroimaging study of pediatric twins. Development and Psychopathology, 20(4), 1161–1175. doi: 10.1017/S0954579408000552 Lenzenweger, M. F. (2013). Endophenotype, intermediate phenotype, biomarker: Definitions, concept comparisons, clarifications. Depression and Anxiety, 30, 185–189. doi: 10.1002/da.22042 Lerner, R. M. (2006). Developmental science, developmental systems, and contemporary theories. In R. M. Lerner (Ed.), Handbook of child psychology: Vol. 1. Theoretical models of human development (6th ed., pp. 1–17). Hoboken, NJ: Wiley. Lerner, R. M. (2009). The positive youth development perspective: Theoretical and empirical bases of a strengths-based approach to adolescent development. In C. R. Snyder & S. J. Lopez (Eds.), Oxford handbook of positive psychology (pp. 149–163). New York, NY: Oxford University Press. Lerner, R. M., & Benson, P. L. (2002). Developmental assets and asset-building communities: Implications for research, policy, and practice. Boston, MA: Kluwer Academic. Lickliter, R. (2013). Biological development: Theoretical approaches, techniques, and key findings. In P. D. Zelazo (Ed.), The Oxford handbook of developmental psychology. Vol. 1: Body and mind (pp. 65–90). New York, NY: Oxford University Press. doi: 10.1093/oxfordhb/9780199958450.013.0004 Lieberman, A. F. (1992). Infant-parent psychotherapy with toddlers. Development and Psychopathology, 4, 559–574. doi: 10.1017/S0954579400004879 Longstaff, P. H. 2009. Managing surprises in complex systems: multidisciplinary perspectives on resilience. Ecology and Society, 14(1): 49 [Online]. Retrieved from http://www.ecologyandsociety.org/vol14/iss1/art49/ Lösel, F., & Farrington, D. P. (2012). Direct and protective and buffering protective factors in the development of youth violence. American Journal of Preventive Medicine, 43, S8–S23. doi: 10.1016/j.amepre.2012.04.029

Lundberg, M., & Wuermli, A. (Eds.). (2012). Children and youth in crisis: Protecting and promoting human development in times of economic shocks. Washington, DC: World Bank. Luo, H., Hu, X., Liu, X., Ma, X., Guo, W., Qiu, C.,…Li, T. (2012). Hair cortisol level as a biomarker for altered hypothalamic-pituitary-adrenal activity in female adolescents with posttraumatic stress disorder after the 2008 Wenchuan earthquake. Biological Psychiatry, 72, 65–69. doi: 10.1016/j.biopsych.2011.12.020 Lupien, S. J., Oullet-Morin, I., Hupbach, A., Tu, M. T., Buss, C., Walker, D.,…McEwen, B. S. (2006). Beyond the stress concept: Allostatic load -a developmental biological and cognitive perspective. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology. Vol. 2: Developmental neuroscience (2nd ed., pp. 578–628). Hoboken, NJ: Wiley. Luthar, S. S. (1991). Vulnerability and resilience: A study of high-risk adolescents. Child Development, 62, 600–616. doi: 10.2307/1131134 Luthar, S. S. (2006). Resilience in development: A synthesis of research across five decades. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology. Vol. 3: Risk, disorder, and adaptation (2nd ed., pp. 739–795). Hoboken, NJ: Wiley. Luthar, S. S., & Cicchetti, D. (2000). The construct of resilience: Implications for intervention and social policy. Development and Psychopathology, 12, 857–885. doi: 10.1017/S0954579400004156 Luthar, S. S., Cicchetti, D., & Becker, B. (2000). The construct of resilience: A critical evaluation and guidelines for future work. Child Development, 71, 543–562. doi: 10.1111/1467–8624.00164 Luthar, S. S., Doernberger, C. H., & Zigler, E. (1993). Resilience is not a unidimensional construct: Insights from a prospective study of inner-city adolescents. Development and Psychopathology, 5, 703–717. doi.org/10.1017/S0954579400006246 Lyons, D. M., & Parker, K. J. (2007). Stress inoculation-induced indications of resilience in monkeys. Journal of Traumatic Stress, 20, 423–433. doi: 10.1002/jts.20265 Maholmes, V. (2014). Fostering resilience and well-being in children and families I poverty: Why hope still matters. New York, NY: Oxford University Press. Mahoney, J. L., Larson, R. W., & Eccles, J. S. (2005). Organized activities as contexts of development: Extracurricular activities, after-school and community programs. Mahwah, NJ: Erlbaum. Manuck, S. B. (2011). Delay discounting covaries with childhood socio-economic status as a function of genetic variation in the dopamine D4 receptor (DRD4). Paper presented at the Society for Research in Child Development, Montreal, Quebec, Canada. Mascolo, M. F., Pollack, R. D., & Fischer, K. W. (1997). Keeping the construction in

development: An epigenetic systems approach. Journal of Constructivist Psychology, 10, 25– 49. doi: 10.1080/10720539708404610 Masten, A. S. (1986). Humor and competence in school-aged children. Child Development, 57, 461–473. doi: 10.1111/1467–8624.ep7266629 Masten, A. S. (1989). Resilience in development: Implications of the study of successful adaptation for developmental psychopathology. In D. Cicchetti (Ed.), The emergence of a discipline: Rochester symposium on developmental psychopathology (Vol. 1, pp. 261–294). Hillsdale, NJ: Erlbaum. Masten, A. S. (1999). Resilience comes of age: Reflections on the past and outlook for the next generation of research. In M. D. Glantz & J. L. Johnson (Eds.), Resilience and development: Positive life adaptations (pp. 281–296). New York, NY: Plenum Press. Masten, A. S. (2001). Ordinary magic: Resilience processes in development. American Psychologist, 56, 227–238. doi: 10.1037/0003–066X.56.3.227 Masten, A. S. (2006). Promoting resilience in development: A general framework for systems of care. In R. J. Flynn, P. Dudding, & J. G. Barber (Eds.), Promoting resilience in child welfare (pp. 3–17). Ottawa, Canada: University of Ottawa Press. Masten, A. S. (2007). Resilience in developing systems: Progress and promise as the fourth wave rises. Development and Psychopathology, 19, 921–930. doi: 10.1017/S0954579407000442 Masten, A. S. (2011). Resilience in children threatened by extreme adversity: Frameworks for research, practice, and translational synergy. Development and Psychopathology, 23, 141– 154. doi: 10.1017/S0954579411000198 Masten, A. S. (2012). Resilience in children: Vintage Rutter and beyond. In A. Slater & P. Quinn (Eds.), Developmental psychology: Revisiting the classic studies (pp. 204–221). London: Sage. Masten, A. S. (2013). Risk and resilience in development. In P. D. Zelazo (Ed.), The Oxford handbook of developmental psychology. Vol. 2: Self and other (pp. 579–607). New York, NY: Oxford University Press. doi: 10.1093/oxfordhb/9780199958474.013.0023 Masten, A. S. (2014a). Global perspectives on resilience in children and youth. Child Development, 85, 6–20. doi: 10.1111/cdev.12205 Masten, A. S. (2014b). Ordinary magic: Resilience in development. New York, NY: Guilford Press. Masten, A. S. (2014c). Promoting the capacity for peace in early childhood: Perspectives from research on resilience in children and families. In J. F. Leckman, C. Panter-Brick, & R. Salah (Eds.), Pathways to peace: The transformative power of families and child development (pp.

251–271). Cambridge MA: MIT Press. Masten, A. S., Best, K. M., & Garmezy, N. (1990). Resilience and development: Contributions from the study of children who overcome adversity. Development and Psychopathology, 2, 425–444. doi: 10.1017/S0954579400005812 Masten, A. S., Burt, K. B., & Coatsworth, J. D. (2006). Competence and psychopathology in development. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology, Vol. 3: Risk, disorder and psychopathology (2nd ed., pp. 696–738). Hoboken, NJ: Wiley. Masten, A. S., & Cicchetti, D. (2010). Developmental cascades. Development and Psychopathology, 22, 491–495. doi: 10.1017/S0954579410000222 Masten, A. S., & Cicchetti, D. (Eds.). (2010a). Developmental cascades: Part 1 [Special issue]. Development and Psychopathology, 22(3), 491–715. doi: 10.1017/S0954579410000222 Masten, A. S., & Cicchetti, D. (Eds.). (2010b). Developmental cascades: Part 2 [Special issue]. Development and Psychopathology, 22(4), 717–983. Masten, A. S., & Cicchetti, D. (Eds.). (2010c). Editorial: Developmental cascades: Part 1 [Special issue]. Development and Psychopathology, 22, 491–495. doi: 10.1017/S0954579410000222 Masten, A. S., & Coatsworth, J. D. (1995). Competence, resilience, and psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology. Vol. 2: Risk, disorder, and adaptation (pp. 715–752). New York, NY: Wiley. Masten, A. S., & Coatsworth, J. D. (1998). The development of competence in favorable and unfavorable environments: Lessons from research on successful children. American Psychologist, 53, 205–220. doi: 10.1037/0003–066X.53.2.205 Masten, A. S., Coatsworth, J. D., Neemann, J., Gest, S. D., Tellegen, A., & Garmezy, N. (1995). The structure and coherence of competence from childhood through adolescence. Child Development, 66, 1635–1659. doi: 10.2307/1131901 Masten, A. S., Cutuli, J. J., Herbers, J. E., Hinz, E., Obradović, J., & Wenzel, A. (2014). Academic risk and resilience in the context of homelessness. Child Development Perspectives, 8, 201–206. doi: 10.1111/cdep.12088 Masten, A. S., & Garmezy, N. (1985). Risk, vulnerability, and protective factors in developmental psychopathology. In B. B. Lahey and A. E. Kazdin (Eds.), Advances in clinical child psychology (Vol. 8, pp. 1–52). New York, NY: Plenum Press. doi: 10.1007/978–1– 4613–9820–2_1 Masten, A. S., Garmezy, N., Tellegen, A., Pellegrini, D. S., Larkin, K., & Larsen, A. (1988). Competence and stress in school children: The moderating effects of individual and family

qualities. Journal of Child Psychology and Psychiatry, 29, 745–764. doi: 10.1111/j.1469– 7610.1988.tb00751.x Masten, A. S., Herbers, J., Cutuli, J., & Lafavor, T. (2008). Promoting competence and resilience in the school context. Professional School Counseling, 12, 76–84. doi: 10.5330/PSC.n.2010–12.76 Masten, A. S., Herbers, J. E., Desjardins, C. D., Cutuli, J. J., McCormick, C. M., Sapienza, J. K.,…Zelazo, P. D. (2012). Executive function skills and school success in young children experiencing homelessness. Educational Researcher, 41, 375–384. doi: 10.3102/0013189X12459883 Masten, A. S., Liebkind, K., & Hernandez, D. J. (Eds.). (2012). Realizing the potential of immigrant youth. New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139094696 Masten, A. S., & Monn, A. R. (2015). Child and family resilience: A call for integrated science, practice, and professional training. Family Relations, 64, 5–21. Masten, A. S., Monn, A. R., & Supkoff, L. M. (2011). Resilience in children and adolescents. In Southwick, S. M., Litz, B., Charney, D. & Friedman, M. J. (Eds.). Resilience: Responding to challenges across the lifespan (pp. 103–119). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511994791.009 Masten, A. S., Morison, P., Pellegrini, D., & Tellegen, A. (1990). Competence under stress: Risk and protective factors. In J. Rolf, A. S. Masten, D. Cicchetti, K. Nuechterlein, & S. Weintraub (Eds.), Risk and protective factors in the development of psychopathology (pp. 236–256). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511752872.015 Masten, A. S., & Narayan, A. J. (2012). Child development in the context of disaster, war and terrorism: Pathways of risk and resilience. Annual Review of Psychology, 63, 227–257. doi: 10.1146/annurev-psych-120710–100356 Masten, A. S., Narayan, A. J., Silverman, W. K., & Osofsky, J. D. (2015). Children in war and disaster. In R. M. Lerner (Ed.), M. H. Bornstein and T. Leventhal (Vol. Eds.), Handbook of child psychology and developmental science. Vol. 4. Ecological settings and processes in developmental systems (7th ed., pp. 704–745). Hoboken, NJ: Wiley. Masten, A. S., & O'Connor, M. J. (1989). Vulnerability, stress, and resilience in the early development of a high risk child. Journal of the American Academy of Child and Adolescent Psychiatry, 28, 274–278. doi: 10.1097/00004583–198903000–00021 Masten, A. S., & Obradovic, J. (2006). Competence and resilience in development. Annals of the New York Academy of Sciences, 1094, 13–27. doi: 10.1196/annals.1376.003 Masten, A. S., & Obradovic, J. (2008). Disaster preparation and recovery: Lessons from

research on resilience in human development. Ecology and Society, 13. Retrieved from www.ecologyandsociety.org/vol13/iss1/art9/ Masten, A. S., & Reed, M.-G. (2002). Resilience in development. In C. R. Snyder & S. J. Lopez (Eds.), Handbook of positive psychology (pp. 74–88). London, UK: Oxford University Press. Masten, A. S., Roisman, G. I., Long, J. D., Burt, K. B., Obradovié, J., Riley, J. R.,…Telegen, A. (2005). Developmental cascades: Linking academic achievement and externalizing and internalizing symptoms over 20 years. Developmental Psychology, 41, 733–746. doi: 10.1037/0012–1649.41.5.733 Masten, A. S., & Tellegen, A. (2012). Resilience in developmental psychopathology: Contributions of the Project Competence Longitudinal Study. Development and Psychopathology, 24, 345–361. doi: 10.1017/S095457941200003X McAdams, D. P., & McLean, K. C. (2013). Narrative identity. Current Directions in Psychological Science, 22, 233–238. doi: 10.1177/0963721413475622 McCormick, C. M., Kuo, S. I.-C., & Masten, A. S. (2011). Developmental tasks across the lifespan. In K. L. Fingerman, C. Berg, T. C. Antonucci, & J. Smith (Eds.), The handbook of lifespan development (pp. 117–140). New York, NY: Springer. McCrory, E., De Brito, S. A., & Viding, E. (2010). Research review: The neurobiology and genetics of maltreatment and adversity. Journal of Child Psychology and Psychiatry, 51, 1079–1095. doi: 10.1111/j.1469–7610.2010.02271.x McEwen, B. S., & Gianaros, P. J. (2011). Stress-and allostasis-induced brain plasticity. Annual Review of Medicine, 62, 431–445. doi: 10.1146/annurev-med-052209–100430 McEwen, B. S., Gray, J. D., & Nasca, C. (2015). Recognizing resilience: Learning from the effects of stress on the brain. Neurobiology of Stress, 1, 1–11. doi: 10.1016/j.ynstr.2014.09.001 McEwen, B. S., & Stellar, E. (1993). Stress and the individual: Mechanisms leading to disease. Archives of Internal Medicine, 153, 2093–2101. doi: 10.1001/archinte.1993.00410180039004 McGloin, J. M., & Widom, C. S. (2001). Resilience among abused and neglected children grown up. Development and Psychopathology, 13, 1021–1038. doi: 10.1017/S095457940100414X McLean, K. C., & Pratt, M. W. (2006). Life's little (and big) lessons: Identity statuses and meaning-making in the turning point narratives of emerging adults. Developmental Psychology, 42, 714–722. doi: 10.1037/0012–1649.42.4.714 Meaney, M. J. (2010). Epigenetics and the biological definition of gene × environment

interactions. Child Development, 81, 41–79. doi: 10.1111/j.1467–8624.2009.01381.x Mill, J. (2011). Epigenetic effects on gene function and their role in mediating geneenvironment interactions. In K. S. Kendler, S. Jaffee, & D. Romer (Eds.), The dynamic genome and mental health (pp. 144–171). New York, NY: Oxford University Press. Miller, G. E., Brody, G. H., Yu, T., & Chen, E. (2014). A family-oriented psychosocial intervention reduces inflammation in low-SES African American youth. Proceedings of the National Academy of Sciences, 111, 11287–11292. doi: 10.1073/pnas.1406578111 Miller, G. E., Chen, E., & Parker, K. J. (2011). Psychological stress in childhood and susceptibility to the chronic diseases of aging: Moving toward a model of behavioral and biological mechanisms. Psychological Bulletin, 137, 959–997. doi: 10.1037/a0024768 Minnesota Department of Health. (2011). Adverse childhood experiences in Minnesota: Findings and recommendations based on the 2011 Minnesota Behavioral Risk Factor Surveillance System. St. Paul, MN: Minnesota Department of Health. http://www.health.state.mn.us/divs/cfh/program/ace/content/document/pdf/acereport.pdf Moffitt, T. E., Caspi, A., & Rutter, M. (2006). Measured gene–environment interactions in psychopathology: Concepts, research strategies, and implications for research, intervention, and public understanding of genetics. Perspectives on Psychological Science, 1, 5–27. doi: 10.1111/j.1745–6916.2006.00002.x Motti-Stefanidi, F. (2014). Immigrant youth adaptation in the Greek school context: A risk and resilience developmental perspective. Child Development Perspectives, 8, 180–185. Murphy, L. (1962). The widening world of childhood: Paths toward mastery. New York, NY: Basic Books. Muthén, B., & Muthén, L. (2000). Integrating person-centered and variable-centered analysis: Growth mixture modeling with latent trajectory classes. Alcoholism: Clinical and Experimental Research, 24, 882–891. doi: 10.1111/j.1530–0277.2000.tb02070.x Nagin, D. S. (1999). Analyzing developmental trajectories: A semiparametric group-based approach. Psychological Methods, 4, 139–157. doi: 10.1037/1082–989X.4.2.139 Nagin, D. S. (2005). Group-based modeling of development. Cambridge, MA: Harvard University Press. Nagin, D. S., & Odgers, C. L. (2010). Group-based trajectory modeling in clinical research. Annual Review of Clinical Psychology, 6, 109–138. doi: 10.1146/annurev.clinpsy.121208.131413 Nagin, D., & Tremblay, R. E. (1999). Trajectories of boys' physical aggression, opposition, and hyperactivity on the path to physically violent and nonviolent juvenile delinquency. Child Development, 70(5), 1181–1196. doi: 10.1111/1467–8624.00086

National Research Council and Institute of Medicine. (2004). Engaging schools: Fostering high school students' motivation to learn. Committee on Increasing High School Students' Engagement and Motivation to Learn. Board on Children, Youth, and Familes. Washington, DC: National Academies Press. National Research Council and Institute of Medicine. (2009). Preventing mental, emotional, and behavioral disorders among young people: Progress and possibilities. Washington, DC: National Academies Press. Nelson, C. A. (1999). Neural plasticity and human development. Current Directions in Psychological Science, 8, 42–45. doi: 10.1111/1467–8721.00010 Nelson, C. A., Fox, N. A., & Zeanah, C. H. (2014). Romania's abandoned children. Cambridge, MA: Harvard University Press. doi: 10.4159/harvard.9780674726079 Nelson, C. A., Zeanah, C. H., Fox, N. A., Marshall, P. J., Smyke, A. T., & Guthrie, D. (2007). Cognitive recovery in socially deprived young children: The Bucharest Early Intervention Project. Science, 318, 1937–1940. doi: 10.1126/science.1143921 Nichols, W. C. (2013). Roads to understanding family resilience: 1920s to the twenty-first century. In D. S. Becvar (Ed.), Handbook of family resilience (pp. 1–16). New York, NY: Springer. doi: 10.1007/978–1–4614–3917–2_1 Nietzsche, F. (2013). Twilight of the idols or How to philosophize with a hammer (D. F. Ferrer Trans.). The Internet Archive (Original work published 1889). Nigg, J., Nikolas, M., Friderici, K., Park, L., & Zucker, R. A. (2007). Genotype and neuropsychological response inhibition as resilience promoters for attentiondeficit/hyperactivity disorder, oppositional defiant disorder, and conduct disorder under conditions of psychosocial adversity. Development and Psychopathology, 19(03), 767–786. doi: 10.1017/S0954579407000387 Norris, F., Stevens, S., Pfefferbaum, B., Wyche, K., & Pfefferbaum, R. (2008). Community resilience as a metaphor, theory, set of capacities, and strategy for disaster readiness. American Journal of Community Psychology, 41, 127–150. doi: 10.1007/s10464–007–9156– 6 Nowakowski, R. S., & Hayes, N. L. (1999). CNS development: An overview. Development and Psychopathology, 11, 395–418. doi: 10.1017/S0954579499002126 Obradović, J., & Boyce, W. T. (2009). Individual differences in behavioral, physiological, and genetic sensitivities to contexts: Implications for development and adaptation. Developmental Neuroscience, 31, 300–308. doi: 10.1159/000216541 Obradović, J., Burt, K. B., & Masten, A. S. (2006). Pathways of adaptation from adolescence to young adulthood: Antecedents and correlates. Annals of the New York Academy of Sciences, 1094, 340–344. doi: 10.1196/annals.1376.046

Obradović, J., Shaffer, A., & Masten, A. S. (2012). Risk and adversity in developmental psychopathology: Progress and future directions. In L. C. Mayes & M. Lewis (Eds.), The Cambridge handbook of environment in human development: A handbook of theory and measurement (pp. 35–37). New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139016827.004 Okada, H., Kuhn, C., Feillet, H., & Bach, J.-F. (2010). The “hygiene hypothesis” for autoimmune and allergic diseases: An update. Clinical and Experimental Immunology, 160, 1–9. doi: 10.1111/j.1365–2249.2010.04139.x Osofsky, H. J., & Osofsky, J. D. (2013). Hurricane Katrina and the Gulf oil spill: Lessons learned. Psychiatric Clinics of North America, 36, 371–383. doi: 10.1016/j.psc.2013.05.009 Overton, W. F. (2013). A new paradigm for developmental science: Relationism and relational- developmental systems. Applied Developmental Science, 17, 94–107. doi: 10.1080/10888691.2013.778717 Panter-Brick, C., Goodman, A., Tol, W., & Eggerman, M. (2011). Mental health and childhood adversities: A longitudinal study in Kabul, Afghanistan. Journal of the American Academy of Child and Adolescent Psychiatry, 50, 349–363. doi: 10.1016/j.jaac.2010.12.001 Panter-Brick, C., & Leckman, J. F. (2013). Editorial commentary: Resilience in child development— interconnected pathways to well-being. Journal of Child Psychology and Psychiatry, 54, 333–336. doi: 10.1111/jcpp.12057 Park, C. L. (2010). Making sense of the meaning literature: An integrative review of meaning making and its effects on adjustment to stressful life events. Psychological Bulletin, 136, 257– 301. doi: 10.1037/a0018301 Park, N. (2011). Military children and families: Strengths and challenges during peace and war. American Psychologist, 66, 65–72. doi: 10.1037/a0021249 Parker, K. J., & Maestripieri, D. (2011). Identifying key features of early stressful experiences that produce stress vulnerability and resilience in primates. Neuroscience & Biobehavioral Reviews, 35, 1466–1483. doi: 10.1016/j.neubiorev.2010.09.003 Patterson, G. R., Forgatch, M. S., & DeGarmo, D. S. (2010). Cascading effects following intervention. Developmental Psychopathology, 22, 941–970. doi: 10.1017/S0954579410000568 Patterson, G. R., Reid, J. B., & Dishion, T. J. (1992). Antisocial boys. Eugene, OR: Castalia. Patterson, J. M. (2002). Integrating family resilience and family stress theory. Journal of Marriage and Family, 64, 349–360. doi: 10.1111/j.1741–3737.2002.00349.x Peterson, B. S., Wang, Z., Horga, G., Warner, V., Rutherford, B., Klahr, K. W.,…Weissman, M. M. (2014). Discriminating risk and resilience endophenotypes from lifetime illness effects in

familial major depressive disorder. JAMA Psychiatry, 71(2), 136–148. doi: 10.1001/jamapsychiatry.2013.4048 Pickreign Stronach, E., Toth, S. L., Rogosch, F. A., & Cicchetti, D. (2013). Preventive interventions and sustained attachment security in maltreated children: A 12-month follow-up of a randomized controlled trial. Development and Psychopathology, 26, 919–930. doi: 10.1017/S0954579413000278 Pluess, M., & Belsky, J. (2013). Vantage sensitivity: Individual differences in response to positive experiences. Psychological Bulletin, 139, 901–916. doi: 10.1037/a0030196 Polanczyk, G., Caspi, A., Williams, B., Price, T. S., Danese, A., Sugden, K.,…Moffitt, T. E. (2009). Protective effects of CRHR1 gene variants on the development of adult depression following childhood maltreatment. Archives of General Psychiatry, 66, 978–985. doi: 10.1001/archgenpsychiatry.2009.114 Porges, S. W. (2011). The polyvagal theory: Neurophysiological foundations of emotions, attachment, communication, and self-regulation. New York, NY: Norton. Prince-Embury, S., & Saklofske, D. H. (Eds.). (2013). Resilience in children, adolescents, and adults: Translating research into practice. New York, NY: Springer. doi: 10.1007/978–1–4614– 4939–3 Prince-Embury, S., & Saklofske, D. (Eds.). (2014). Resilience interventions for youth in diverse populations. New York, NY: Springer. doi: 10.1007/978–1–4939–0542–3 Qouta, S., Punamäki, R., & El Sarraj, E. (2008). Child development and family mental health in war and military violence: The Palestinian experience. International Journal of Behavioral Development, 32, 310–321. doi: 10.1177/0165025408090973 Ramchandahl, S., Bhattacharya, S. K., Cervoni, N., & Szyf, M. (1999). DNA methylation is a reversible biological signal. Proceedings of the National Academy of Sciences of the United States of America, 96(11), 6107–6112. doi: 10.1073/pnas.96.11.6107 Raver, C. C. (2012). Low-income children's self-regulation in the classroom: Scientific inquiry for social change. American Psychologist, 67, 681–689. doi: 10.1037/a0030085 Raver, C. C., Li-Grining, C., Bub, K., Jones, S. M., Zhai, F., & Pressler, E. (2011). CSRP's impact on low-income preschoolers' preacademic skills: Self-regulation as a mediating mechanism. Child Development, 82, 362–378. doi: 10.1111/j.1467–8624.2010.01561.x Repetti, R. L., Robles, T. F., & Reynolds, B. (2011). Allostatic processes in the family. Development and Psychopathology, 23, 921–938. doi: 10.1017/S095457941100040X Reynolds, A. J., Temple, J. A., White, B. A. B., Ou, S.-R., & Robertson, D. L. (2011). Age 26 cost-benefit analysis of the Child–Parent Center early education program. Child Development, 82, 379–404. doi: 10.1111/j.1467–8624.2010.01563.x

Roth, T. L. (2013). Epigenetic mechanisms in vulnerability and resilience to psychopathology: Advances, challenges, and future promises of a new field. Development and Psychopathology, 25, 1279–1291. doi: 10.1017/S0954579413000618 Rothbart, M. K. (2011). Becoming who we are: Temperament and personality in development. New York, NY: Guilford Press. Rowe, J. W., & Kahn, R. L. (Eds.). (1998). Successful aging. New York, NY: Pantheon Books. Rutter, M. (1979). Protective factors in children's responses to stress and disadvantage. In M. W. Kent & J. E. Rolf (Eds.), Primary prevention of psychopathology: Vol. 3. Social competence in children (pp. 49–74). Hanover, NH: University Press of New England. Rutter, M. (1987). Psychosocial resilience and protective mechanisms. American Journal of Orthopsychiatry, 57, 316–331. doi: 10.1111/j.1939–0025.1987.tb03541.x Rutter, M. (1999). Resilience concepts and findings: Implications for family therapy. Journal of Family Therapy, 21, 119–144. doi: 10.1111/1467–6427.00108 Rutter, M. (2006). Implications of resilience concepts for scientific understanding. Annals of the New York Academy of Sciences, 1094, 1–12. doi: 10.1196/annals.1376.002 Rutter, M. (2012). Resilience as a dynamic concept. Development and Psychopathology, 24(2), 335. doi: 10.1017/S0954579412000028 Rutter, M. (2013). Annual research review: Resilience—clinical implications. Journal of Child Psychology and Psychiatry, 54, 474–487. doi: 10.1111/j.1469–7610.2012.02615.x Rutter, M., & Garmezy, N. (1983). Developmental psychopathology. In E. M. Hetherington (Ed.), Mussen's handbook of child psychology: Vol. 4 Socialization, personality, and social development (4th ed., pp. 775–911). New York, NY: Wiley. Rutter, M., Kim-Cohen, J., & Maughan, B. (2006). Continuities and discontinuities in psychopathology between childhood and adult life. Journal of Child Psychology and Psychiatry, 47, 276–295. doi: 10.1111/j.1469–7610.2006.01614.x Rutter, M., & Maughan, B. (2002). School effectiveness findings 1979–2002. Journal of School Psychology, 40, 451–475. doi: 10.1016/S0022–4405(02)00124–3 Rutter, M., Moffitt, T. E., & Caspi, A. (2006). Gene–environment interplay and psychopathology: multiple varieties but real effects. Journal of Child Psychology and Psychiatry, 47, 226–261. doi: 10.1111/j.1469–7610.2005.01557.x Rutter, M., Sonuga-Barke, E. J., & Castle, J. (2010). Investigating the impact of early institutional deprivation on development: Background and research strategy of the English and Romanian Adoptees (ERA) study. Monographs of the Society for Research in Child Development, 75, 1–20. doi: 10.1111/j.1540–5834.2010.00548.x

Rutter, M., & Sroufe, L. A. (2000). Developmental psychopathology: Concepts and challenges. Development and Psychopathology, 12(3), 265–296. doi: 10.1017/s0954579400003023 Ryan, R. M., & Deci, E. L. (2000). Self-determination theory and the facilitation of intrinsic motivation, social development, and well-being. American Psychologist, 55, 68–78. doi: 10.1037/0003–066X.55.1.68 Sameroff, A. J. (2000). Developmental systems and psychopathology. Development and Psychopathology, 12(3), 297–312. doi: 10.1017/s0954579400003035 Sameroff, A. J. (2006). Identifying risk and protective factors for healthy child development. In A. Clark-Stewart & J. Dunn (Eds.), Families count: Effects on child and adolescent development (pp. 53–76). Cambridge, UK: Cambridge University Press. Sameroff, A. J. (2010). A unified theory of development: A dialectic integration of nature and nurture. Child Development, 81, 6–22. doi: 10.1111/j.1467–8624.2009.01378.x Sameroff, A. J., & Chandler, M. J. (1975). Reproductive risk and the continuum of caretaking casualty. In F. D. Horowitz, E. M. Hethterington, S. Scarr-Salapatek, & G. M. Siegel (Eds.), Review of child development research (Vol. 4, pp. 187–243). Chicago, IL: University of Chicago Press. Sameroff, A. J., & Seifer, R. (1983). Familial risk and child competence. Child Development, 54, 1254–1268. doi: 10.2307/1129680 Sameroff, A. J., Seifer, R., & Bartko, W. T. (1997). Environmental perspectives on adaptation during childhood and adolescence. In S. S. Luthar, J. A. Burack, D. Cicchetti, & J. R. Weisz (Eds.), Developmental psychopathology: Perspectives on adjustment, risk, and disorder (pp. 507–526). New York, NY: Cambridge University Press. Sameroff, A. J., Seifer, R., Barocas, R., Zax, M., & Greenspan, S. (1987). Intelligence quotient scores of 4-year-old children: Social–environmental risk factors. Pediatrics, 79, 343–350. Sandler, I., Schoenfelder, E., Wolchik, S., & MacKinnon, D. (2011). Long-term impact of prevention programs to promote effective parenting: Lasting effects but uncertain processes. Annual Review of Psychology, 62, 299–329. doi: 10.1146/annurev.psych.121208.131619 Sapienza, J. K., & Masten, A. S. (2011). Understanding and promoting resilience in children and youth. Current Opinion in Psychiatry, 24, 267–273. doi: 10.1097/YCO.0b013e32834776a8 Scales, P. C., Benson, P. L., Leffert, N., & Blyth, D. A. (2000). Contribution of developmental assets to the prediction of thriving among adolescents. Applied Developmental Science, 4, 27– 46. doi: 10.1207/S1532480XADS0401_3 Scales, P. C., Benson, P. L., Roehlkepartain, E. C., Sesma, A., Jr., & van Dulmen, M. (2006). The role of developmental assets in predicting academic achievement: A longitudinal study.

Journal of Adolescence, 29I, 691–708. doi: 10.1016/j.adolescence.2005.09.001 Schoon, I. (2006). Risk and resilience: Adaptations in changing times. New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511490132 Serafica, F. C., & Vargas, L. A. (2006). Cultural diversity in the development of child psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Vol. 1. Theory and method (2nd ed., pp. 588–626). Hoboken, NJ: Wiley. Sesma, A., Mannes, M., & Scales, P. C. (2013). Positive adaptation, resilience and the developmental assets framework. In S. Goldstein & R. B. Brooks (Eds.), Handbook of resilience in children (pp. 427–442). New York, NY: Springer. doi: 10.1007/978–1–4614– 3661–4_25 Sheinkopf, S. J., Lagasse, L. L., Lester, B. M., Liu, J., Seifer, R., Bauer, C. R.,…Das, A. (2007). Vagal tone as a resilience factor in children with prenatal cocaine exposure. Development and Psychopathology, 19(3), 649–673. doi: 10.1017/S0954579407000338 Shonkoff, J. P. (2011). Protecting brains, not simply stimulating minds. Science, 333, 982–983. doi: 10.1126/science.1206014 Singer, W. (1995). Development and plasticity of cortical processing architectures. Science, 270, 758–764. doi: 10.1126/science.270.5237.758 Sirin, S. R., & Gupta, T. (2012). Muslim, American, and immigrant: Integration despite challenges. In A. S. Masten, K. Liebkind, & D. J. Hernandez (Eds.), Realizing the potential of immigrant youth (pp. 253–279). New York, NY: Cambridge University Press. doi: 10.1017/CBO9781139094696.013 Small, S., & Memmo, M. (2004). Contemporary models of youth development and problem prevention: Toward an integration of terms, concepts, and models. Family Relations, 53, 3–11. doi: 10.1111/j.1741–3729.2004.00002.x Smart, E., & Stewart, C. (2013). My story. New York, NY: St. Martin's Press. Snyder, C. R., Hoza, B., Pelham, W. E., Rapoff, M., Ware, L., Danovsky, M.,…Stahl, K. J. (1997). The development and validation of the children's hope scale. Journal of Pediatric Psychology, 22, 399–421. doi: 10.1093/jpepsy/22.3.399 Southwick, S. M., Bonanno, G. A., Masten, A. S., Panter-Brick, C., & Yehuda, R. (2014). Resilience definitions theory, and challenges: Interdisciplinary perspectives. European Journal of Psychotraumatology, 5, 25338 (1–14). doi: 10.3402/ejpt.v5.25338 Southwick, S. M., Litz, B. T., Charney, D., & Friedman, M. J. (Eds.). (2011). Resilience and mental health: Challenges across the lifespan. New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511994791 Sroufe, L. A. (1979). The coherence of individual development: Early care, attachment, and

subsequent developmental issues. American Psychologist, 34, 834–841. doi: 10.1037/0003– 066X.34.10.834 Sroufe, L. A. (1983). Infant-caregiver attachment patterns of adaptation in preschool: The roots of maladaptation and competence. In M. Perlmutter (Ed.), Minnesota symposium on child psychology, Vol. 16 (pp. 41–83). Hillsdale, NJ: Erlbaum. Sroufe, L. A. (1989). Pathways to adaptation and maladaptation: Psychopathology as developmental deviation. In D. Cicchetti (Ed.), Rochester symposia on developmental psychopathology, Vol. 1 (pp. 13–40). Hillsdale, NJ: Erlbaum. Sroufe, L. A. (2007). The place of development in developmental psychopathology. In A. Masten (Ed.), Multilevel dynamics in developmental psychopathology: Pathways to the future: The Minnesota symposia on child psychology, Vol. 34 (pp. 285–299). Mahwah, NJ: Erlbaum. Sroufe, L. A. (2013). The promise of developmental psychopathology: Past and present. Development and Psychopathology, 25, 1215–1224. doi: 10.1017/S0954579413000576 Sroufe, L. A., Carlson, E. A., Levy, A. K., & Egeland, B. (1999). Implications of attachment theory for developmental psychopathology. Development and Psychopathology, 11, 1–13. Sroufe, L. A., Egeland, B., Carlson, E. A., Collins, W. A. (2005). The development of the person: The Minnesota study of risk and adaptation from birth to adulthood. New York, NY: Guilford Press. Sroufe, L. A., Egeland, B., & Kreutzer, T. (1990). The fate of early experience following developmental change: Longitudinal approaches to individual adaptation in childhood. Child Development, 61, 1363–1373. doi: 10.2307/1130748 Sroufe, L. A., & Rutter, M. (1984). The domain of developmental psychopathology. Child Development, 55, 17–29. doi: 10.2307/1129832 Steinberg, L. (2008). A social neuroscience perspective on adolescent risk-taking. Developmental Review, 28, 78–106. doi: 10.1016/j.dr.2007.08.002 Stevens, H. E., Leckman, J. F., Coplan, J. D., & Suomi, S. J. (2009). Risk and resilience: Early manipulation of macaque social experience and persistent behavioral and neurophysiological outcomes. Journal of the American Academy of Child and Adolescent Psychiatry, 48, 114– 127. doi: 10.1097/CHI.0b013e318193064c Suomi, S. J. (2006). Risk, resilience, and gene × environment interactions in rhesus monkeys. Annals of the New York Academy of Sciences, 1094, 52–62. doi: 10.1196/annals.1376.006 Suomi, S. J. (2011). Risk, resilience, and gene–environment interplay in primates. Journal of the Canadian Academy of Child and Adolescent Psychiatry, 20, 289–297. Szanton, S. L., & Gill, J. M. (2010). Facilitating resilience using a society-tocells framework:

A theory of nursing essentials applied to research and practice. Advances in Nursing Science, 33, 329–343. doi: 10.1097/ANS.0b013e3181fb2ea2 Szyf, M. & Bick, J. (2013). DNA methylation: A mechanism for embedding early life experiences in the genome. Child Development, 84(1), 49–57. doi: 10.1111/j.1467– 8624.2012.01793.x Tarullo, A. R., & Gunnar, M. R. (2006). Child maltreatment and the developing HPA axis. Hormones and Behavior, 50, 632–639. doi: 10.1016/j.yhbeh.2006.06.010 Teicher, M. H., Tomoda, A., & Andersen, S. L. (2006). Neurobiological consequences of early stress and childhood maltreatment: Are results from human and animal studies comparable? Annals of the New York Academy of Sciences, 1071, 313–323. doi: 10.1196/annals.1364.024 Temple, J. A., & Reynolds, A. J. (2007). Benefits and costs of investments in preschool education: Evidence from the Child–Parent Centers and related programs. Economics of Education Review, 26(1), 126–144. doi: 10.1016/j.econedurev.2005.11.004 Theron, L. C., Theron, A. M. C., & Malindi, M. J. (2013). Toward an African definition of resilience: A rural South African community's view of resilient Basotho youth. Journal of Black Psychology, 39, 63–87. doi: 10.1177/0095798412454675 Thompson, R. A. (2000). The legacy of early attachments. Child Development, 71, 145–152. doi: 10.1111/1467–8624.00128 Toth, S. L., & Gravener, J. (2012). Bridging research and practice: Relational interventions for maltreated children. Child and Adolescent Mental Health, 17, 131–138. doi: 10.1111/j.1475– 3588.2011.00638.x Toth, S. L., & Manley, J. T. (2011). Bridging research and practice: Challenges and successes in implementing evidence-based preventive intervention strategies for child maltreatment. Child Abuse and Neglect, 35, 633–636. doi: 10.1111/1467–8624.00128 Toth, S. L., Rogosch, F. A., Manly, J. T., & Cicchetti, D. (2006). The efficacy of toddler-parent psychotherapy to reorganize attachment in the young offspring of mothers with major depressive disorder: A randomized preventive trial. Journal of Consulting and Clinical Psychology, 74(6), 1006–1016. doi: 10.1037/0022–006X.74.6.1006 Trickett, P. K., Gordis, E., Granger, D. A., & Susman, E. J. (2008). Salivary alpha amylase cortisol asymmetry in maltreated youth. Hormones and Behavior, 53, 96–103. doi: 10.1016/j.yhbeh.2007.09.002 Tronick, E., & Beeghly, M. (2011). Infants' meaning-making and the development of mental health problems. American Psychologist, 66, 107–119. doi: 10.1037/a0021631 Tyrka, A. R., Price, L. H., Marsit, C., Walters, O. C., & Carpenter, L. L. (2012). Childhood adversity and epigenetic modulation of the leukocyte glucocorticoid receptor: Preliminary

findings in healthy adults. PLoS ONE, 7(1). e30148 doi: 10.1371/journal.pone.0030148 Ungar, M. (2008). Resilience across cultures. British Journal of Social Work, 38, 218–235. doi: 10.1093/bjsw/bcl343 Ungar, M. (Ed.). (2012). The social ecology of resilience: A handbook of theory and practice. New York, NY: Springer. doi: 10.1007/978–1–4614–0586–3 Ungar, M., Ghazinour, M., & Richter, J. (2013). What is resilience within the social ecology of human development? Journal of Child Psychology and Psychiatry, 54, 348–366. doi: 10.1111/jcpp.12025 Ungar, M., & Liebenberg, L. (2005). The International Resilience Project. In M. Ungar (Ed.), Handbook for working with children and youth: Pathways to resilience across cultures and contexts (pp. 211–225). Thousand Oaks, CA: Sage. doi: 10.4135/9781412976312.n13 UNICEF. (2011). Humanitarian action for children: Building resilience. New York, NY: United Nations Children's Fund. Valle, M. F., Huebner, E. S., & Suldo, S. M. (2006). An analysis of hope as a psychological strength. Journal of School Psychology, 44, 393–406. doi: 10.1016/j.jsp.2006.03.005 van der Werff, S. J., Pannekoek, J. N., Veer, I. M., van Tol, M. J., Aleman, A., Veltman, D. J., …van der Wee, N. J. (2013). Resting-state functional connectivity in adults with childhood emotional maltreatment. Psychological Medicine, 43(9), 1825–1836. doi: 10.1017/S0033291712002942 van IJzendoorn, M. H., & Bakermans-Kranenburg, M. J. (2015). Genetic differential susceptibility on trial: Meta-analytic support from randomized controlled experiments. Development and Psychopathology, 27, 152–162. doi: 10.1017/S0954579414001369 van Praag, H., Kempermann, G., & Gage, F. H. (2000). Neural consequences of environmental enrichment. Nature Reviews Neuroscience, 1, 191–198. doi: 10.1038/35044558 Vanderwert, R. E., Marshall, P. J., Nelson, C. A., Zeanah, Ch. H., & Fox, N. A. (2010). Timing of intervention affects brain electrical activity in children exposed to severe psychosocial neglect. PLoS ONE, 5(7):e11415. doi: 10.1371/journal.pone.0011415. Volkow, N. D., Wang, G.-J., Fowler, J. S., Tomasi, D., & Telang, F. (2011). Addiction: Beyond dopamine reward circuitry. Proceedings of the National Academy of Sciences, 108, 15037– 15042. doi: 10.1073/pnas.1010654108 von Bertalanffy, L. (1968). General system theory: Foundation, development, application. New York, NY: Braziller. von Mutius, E., & Radon, K. (2008). Living on a farm: Impact on asthma induction and clinical course. Immunology and Allergy Clinics of North America, 28, 631–647. doi: 10.1016/j.iac.2008.03.010

Wachs, T. and A. Rahman. (2013). The nature and impact of risk and protective influences on children's development in low and middle income countries. In P. R. Britto, P. L. Engle, & C. M. Super (Eds.), The handbook of early childhood development research and its impact on global policy (pp. 85–122). New York, NY: Oxford University Press. doi: 10.1093/acprof:oso/9780199922994.003.0005 Waddington, C. H. (1957/2014). The strategy of the genes: A discussion of some aspects of theoretical biology. New York, NY: Routledge. Walsh, F. (2006). Strengthening family resilience (2nd ed.). New York, NY: Guilford Press. Walsh, F. (Ed.). (2011). Family processes: Growing diversity and complexity. New York, NY: Guilford Press. Walsh, F. (2013). Community-based practice applications of a family resilience framework. In D. S. Becvar (Ed.), Handbook of family resilience (pp. 65–82). New York, NY: Springer. doi: 10.1007/978–1–4614–3917–2_5 Watt, N. F., Anthony, E. J., Wynne, L. C., & Rolf, J. E. (Eds.). (1984). Children at risk for schizophrenia: A longitudinal perspective. New York, NY: Cambridge University Press. Werner, E. E. (2000). Through the eyes of innocents: Children witness World War II. Boulder, CO: Westview Press. Werner, E. E., & Smith, R. S. (1982). Vulnerable but invincible: A longitudinal study of resilient children and youth. New York, NY: McGraw-Hill. Werner, E. E., & Smith, R. S. (1992). Overcoming the odds: High risk children from birth to adulthood. Ithaca, NY: Cornell University Press. Werner, E. E., & Smith, R. S. (2001). Journeys from childhood to midlife: Risk, resilience, and recovery. Ithaca, NY: Cornell University Press. White, R. W. (1959). Motivation reconsidered: The concept of competence. Psychological Review, 66, 297–333. doi: 10.1037/h0040934 Winslow, E. B., Sandler, I. N., Wolchik, S. A., & Carr, C. (2013). Building resilience in all children: A public health approach. In S. Goldstein & R. B. Brooks (Eds.), Handbook of resilience in children (pp. 459–480). New York, NY: Springer. doi: 10.1007/978–1–4614– 3661–4_27 Wolmer, L., Hamiel, D., & Laor, N. (2011). Preventing children's posttraumatic stress after disaster with teacher-based intervention: A controlled study. Journal of the American Academy of Child and Adolescent Psychiatry, 50, 340–348.e2. doi: 10.1016/j.jaac.2011.01.002 Wright, M. O'D., Masten, A. S., & Narayan, A. J. (2013). Resilience processes in

development: Four waves of research on positive adaptation in the context of adversity. In S. Goldstein & R. B. Brooks, Handbook of resilience in children (2nd ed., pp. 15–37). New York, NY: Springer. doi: 10.1007/978–1–4614–3661–4_2 Yang, B. Z. Y., Zovehang, H., Ge, W., Weder, N., Douglas-Palumberi, H, Perepletchikova, F., …Kaufman, J. (2013). Child abuse and epigenetic mechanisms of disease risk. American Journal of Preventive Medicine, 44, 101–107. doi: 10.1016/j.amepre.2012.10.012 Zeanah, C. H., Nelson, C. A., Fox, N. A., Smyke, A. T., Marshall, P., Parker, S. W., & Koga, S. (2003). Designing research to study the effects of institutionalization on brain and behavioral development: The Bucharest early intervention project. Development and Psychopathology, 15, 885–907. doi: 10.1017/S0954579403000452 Zelazo, P. D. (2013). Developmental psychology: A new synthesis. In P. D. Zelazo (Ed.), The Oxford handbook of developmental psychology. Vol. 1: Body and mind (pp. 3–12). New York, NY: Oxford University Press. Zelazo, P. D., & Carlson, S. M. (2012). Hot and cool executive function in childhood and adolescence: Development and plasticity. Child Development Perspectives, 6, 354–360. Zelazo, P. D., & Lyons, K. E. (2011). Mindfulness training in childhood. Human Development, 54, 61–65. doi: 10.1159/000327548 Zimmerman, M. A., Stoddard, S. A., Eisman, A. B., Caldwell, C. H., Aiyer, S. M., & Miller, A. (2013). Adolescent resilience: Promotive factors that inform prevention. Child Development Perspectives, 7, 215–220. doi: 10.1111/cdep.12042

Chapter 7 Vulnerability and Resiliency of African American Youth: Revelations and Challenges to Theory and Research Margaret Beale Spencer and Dena Phillips Swanson Until lions have their own historians, tales of the hunt shall always glorify the hunters who tell it. —Ewe-mina (Benin, Ghana, and Togo) Proverb

Everybody is a genius. But if you judge a fish by its ability to climb a tree, it will live its whole life believing that it is stupid. —Albert Einstein

HISTORICAL AND CONTEXTUAL FRAMING OF VULNERABILITY AND RESILIENCY Contextual Background: America's Race Salience Dilemma Perspective and Significance HUMAN VULNERABILITY AND RESILIENCY An Inclusive Perspective: Phenomenological Variant Of Ecological Systems Theory Historical Placement Salient Perspectives Required to Appreciate Challenges to Resiliency New Concepts and “Traditional” Assumptions: Thriving and Development Black Identity and Self-Esteem Research Acknowledging Colorism METHODOLOGICAL CONSIDERATIONS Conceptual Issues Procedural Issues TRANSLATIONAL PRACTICES Work Connections Restorative Practices and Principles CONCLUSION: REFRAMING THE FUTURE REFERENCES For the nation's citizens of color, 150 years following President Abraham Lincoln's signing of

the Emancipation Proclamation and notwithstanding Barack Obama's election during the interim as the first African American and forty-fourth President of the United States of America, unequal access to supports and opportunities persists. The relationship between the presence of accessible supports and the level of challenge confronted (i.e., given myriad risks), considered together, potentially makes the fact of individuals' unavoidable human vulnerability status both more complex and potentially critical for pursuing (traditional) human development tasks (see Havighurst, 1953). The specific level and character of vulnerability is necessarily fluid, context linked and, thus, demonstrates plasticity. Although this fact is uncomfortable for some to acknowledge, and infrequently evidenced in the design of research and in its interpretations, at any given time irrespective of gender, race, ethnicity, skin color, and place of birth, all humans' daily experiences represent some level of vulnerability.

Historical and Contextual Framing of Vulnerability and Resiliency Individuals' lives represent ever-changing exposure to assets and resources as well as experiences of risks and challenges (see Spencer 1995, 2006; Spencer, Harpalani, Cassidy, et al., 2006). For America's minorities, the insufficiency and uncontrollability of assets and the challenges confronted have much to do with their identifiability as “the other” and persistent situations of high risk. Multiple studies explicitly express the role of resiliency among African American youth and also acknowledge the need for them to cross “cultural boundaries” and to use achievement as a way of coping with stress linked with socially constructed conditions (Phelan, Davidson, & Cao, 1991; Pollard, 1989; Siddle-Walker, 1996). From the program of research conducted by Pollard (1989), the character of findings resulted in recommendations targeted to educators which suggested that they encourage students to have high expectations and reward youth for their achievement efforts. During the same time, Werner's (1989) study reported longitudinal findings on the developmental trajectory of children that had experienced high-risk pre- and perinatal births. By the mid-1990s, several studies were examining resiliency for African American youth. Since that time, these studies have evolved with increasing attention to theoretical orientations, considerations of environmental influences inclusive of, but beyond that of the family, and nuanced methodological and statistical approaches for analyzing complex phenomena. Despite these advances and recommendations, for many African American youth, outcomes have declined in education, in physical health, in mental health, and in more general well-being. To address these contradictory trends, this chapter provides a sociohistorical framing in reviewing research that has informed the field, with specific attention on African American youth vulnerability and resilience.

Contextual Background: America's Race Salience Dilemma Obtaining achieved good outcomes while facing significant obstacles is generally viewed as demonstrations of resiliency (see Spencer & Swanson, 2013). Accordingly, human

vulnerability and resiliency are inherently intertwined; the former represent our human experience of normative and nonnormative challenges and supports, and the latter indicate successful outcomes albeit achieved under conditions of significant risk. Of course, efforts to achieve resiliency for self and others is not without dire consequences, as Sherman James and colleagues argued (e.g., see James, Keenan, Strogatz, Browning, & Garrett, 1992). Their patterned findings as epidemiological reports suggested an inverse relationship between socioeconomic status (SES) and hypertension. This paradoxical relationship between effort and health-challenging outcomes is associated with prolonged exposure to stress and is called John Henryism; their characterization follows. John Henryism is a strong behavioral disposition to cope in an active, effortful manner with the psychosocial stressors of everyday life. The latter include financial pressures, job demands, parenting, and managing interpersonal conflict, among other things. Persons manifesting the John Henryism coping style believe that daily stressors of this kind can be effectively managed (and high levels of psychological stress thus avoided) if they work hard and are determined to succeed. (p. 15) Implicit in the James et al. (1992) analysis is that adult daily stressors may compromise the adequate provision of supports needed for carrying out developmental tasks due to racial gap differences for hypertension, diabetes, and oft-deadly stroke outcomes. Thus, adults' efforts to successfully meet developmental tasks relevant for themselves, their children, and others for whom they have responsibility are complex (see Havighurst, 1953). Life course efforts and goals designed to meet developmental tasks are not independent of the noted effort described by James et al. Also relevant is the widely disseminated perspective by Swedish economist sociologist Gunnar Myrdal (1944), who posited a particular European perspective. In fact, virtually 50 years before the James et al. (1992) theory, Myrdal framed racial gap differences as an American dilemma and linked it to group membership. With a focus parallel to that of James et al., Myrdal also noted the significant physiologically adaptive costs for minority-status individuals. The analysis of African American historian Ralph Ellison (1944), a contemporary of Myrdal, posited that the process of living in a democracy and adhering to inherent democratic ethos represents the actual dilemma; thus, with a focus on the historical context, Ellison posited that the considerable physical and psychological costs hypothesized are related to the American Creed, which promises success with hard work. Our synthesis of the points of view salient for African American youth resilience and vulnerability implies that the extreme stress-creating dilemmas experienced by socializing adults and youth cannot be isolated from their efforts to tackle normative, stage-specific traditional lifelong developmental requirements. Tasks relevant to the life course call for a variety of supports and accessible assets (i.e., physical, social, economic, and historical)—both for youth and for socializing adults. Research shows that African American parents embrace both American and African-based values and endeavor to instill both value systems in their children (Spencer, 1990b). Parents' values and history of sociocultural experiences with discrimination affect the socialization

strategies instituted (D. Hughes & Johnson, 2001; McHale et al., 2006). Parents value honesty, academic success, and family responsibility and teach these values to their children. In addition, adults are likely to embrace culturally distinct values, which include among others kin networks, respect for the elderly, and mutual cooperation and sharing (S. A. Hill, 2001; Murry, Berkel, Brody, Miller, & Chen, 2009). An underreported perspective is that parenting adults emphasize humanistic values over more ethnic specific parenting practices and values (Marshall, 1995). African American parents want to raise children with values and expectations common with all parents. At the same time, however, they recognize that although they raise their children to treat others with respect, they will not always encounter respect from others or have access to the same resources and supports as others. Accordingly, adults of color and particularly African American socializing agents have the additional and unique task of communicating to youth that the American Creed may not apply to their experiences. Relative to John Henryism (see James et al., 1992), also critically important to socialization effort and character is the reality that minority adults are required to communicate the socialization-relevant message noted without jeopardizing youth motivation, effort, and hope. Relatedly, education has historical value linked to the belief that knowledge represents power and serves as a mechanism for obtaining economic stability. Parental experience with controlled access and lack of receptivity in schools has strained their own efforts for engaged involvement (Hanlon, Simon, O'Grady, Carswell, & Callaman, 2009). Given the wealth of research supporting the impact of parental involvement on positive academic outcomes for students, Black parents' perceived lack of involvement, as inferred by school personnel and defined largely as responding to calls from school and appearing for teacher conferences, is characterized as disinterest in the education of their children (Deslander & Bertrand, 2005; Gutman & McLloyd, 2000). The distancing of parents from the school also compromises the parents' access to information and informal factors relevant to resources. D'Agostino, Hedges, and Borman (2001) suggest that schools, specifically Title I schools designated as schools with a high enrollment of low-income students, create comprehensive procedures to engage parents who may also be confronting a myriad of unacknowledged circumstances critical to the socialization and support of students of color. Such procedures continue to be minimally implemented, creating limited accessibility for supportive educational opportunities for students and their families. Patterns of asset accessibility (or not) may be connected with group membership and assumptions concerning social status that suggest inherent group value (e.g., given race, faith group, immigration status, national origin, skin color, and ethnicity). Unfortunately, the historically linked and broadly diverse challenges confronted by individuals and communities of color remain largely invisible, unacknowledged, and, in fact, minimally addressed by the social science and policy communities. Silencing discussions of disparity etiologies from a human vulnerability perspective, in fact, actually communicates the salience and dilemma of the persistent uneven resource distribution between groups of North Americans residing in the 50 states. In many ways, the situation is well illustrated by the African proverb cited at the beginning of this chapter. Specifically, the silencing of devalued points of view, particularly evident when pertaining to race/ethnicity comparisons and suggested shortcomings, matters; the

situation determines whose “voice” is used for framing the experiences associated with one's racial group and thus the character of policies intended to be supportive. The situation contributes to the American dilemma as described by Myrdal (1944). Ellison's (1944) review of the work, however, noted that the continuing social ambiguity of the nation specifically concerning Blacks had more to do with the nation's Whites experiencing a clash between the American creed and what were actually everyday anti-American practices. The lack of resolution of the conundrum described is not unexpected and influences the interpretation and framing of reports of patterned “gap” findings. The patterns of broadly disseminated between-group racial differences and outcomes, in fact, are merely further stereotypes of already maligned groups. Specifically, and as a case in point, the problematic descriptions and general pathologizing of African American males' life course experiences demonstrate the conceptual dilemma described. The highly negative reporting and limited historical conceptualizing of the lives of African American males is evident as early as the preschool years and persists across the life course. For minority-status individuals more generally, the framing of gap findings frequently adds to active “problematizing” of lives rather than seeking and achieving insights as to contributing forces and persistent social conditions; thus, the dilemma suggests intergenerational significance. The work by Gibbs (1988) on Black males as an endangered species was designed to shock the nation into recognizing the impact negative and persistent societal conditions were having on their life trajectory. The emphasis was designed not to problematize but to create urgency for mitigating action. Rather than the latter, further stigmatizing occurred, creating heightened stress for caring adults who wished to protect youth from experiences that undermine their opportunities. Irrespective of the success level achieved, the stress associated with the experience of group stereotyping, oppressive situations, socially limiting conditions, and adverse health consequences of adults as a function of the proffered efforts required, in fact, persistent and subsequently remain problematic. For example, the literacy rates for Black boys at third grade are framed and disseminated broadly as the critical statistic used for determining the number of prisons requiring construction (see Neal & Rick, 2014). Similarly, the frequently disseminated statistical descriptions that note the existence of more Black males in prison than in college could be framed differently. That is, given that prenatal experiences and neonatal status of Black boys are not unlike those of other groups, literacy rate differences by third grade could be framed in ways which (a) question disparate and persistent contextual experiences, (b) consider accessible supports, and (c) evaluate opportunities had (or not) within the first 8 years of life. For example, a curious and consistent finding is that the status of Black boys is frequently linked to family SES. What is it about a low-resource socialization context that increases the high vulnerability of Black boys over that of girls (e.g., see Swanson, Cunningham, & Spencer 2003)? Given James et al.'s (1992) speculations concerning the dilemma of John Henryism and the racial differences in mortality and morbidity rates, it can be inferred that the social and contextual character of the life course suggests a dire prognosis for males of African descent. Accordingly, questions might inquire about: (a) the extraordinarily untoward social, emotional, and achievement-relevant experiences of Black boys in the first 8 years of life (e.g.,

achievement gaps, “special needs” labeling patterns, too few fathering figures, and contextlinked coping responses to environmental experiences that label them as dangerous); (b) the character of and supportive capacities available in low-resourced communities (e.g., disproportionate contact with police for male adults and youth); and (c) compromised adult health status and socialization capacity to meet developmental task requirements. As indicated, the prenatal developmental status and perinatal experiences had by Black boys do not suggest anticipated gaps in cognitive, affective, or physiological differences by third grade. However, the compromising economic and stress relevant experiences had in the first 8 to 10 years appear impactful and significant in enough character for offsetting inferred developmental advantages evident at birth (i.e., given the absence of early gaps). Reports indicate neither physiological nor cognitive statuses of Black boys at birth that would indicate untoward middle childhood and early adulthood disparate outcomes (e.g., disproportionate Individualized Education Program assignment in school, contact with the criminal justice system, and more adult “prison time” convictions than college attendance when compared proportionally with similarly situated White youth). As suggested by the epigraph by Albert Einstein, erroneous criteria or descriptors may be used for assessing youth of color and particularly Black boys. Frequently, undifferentiating explanatory labeling used for describing gap differences, such as “structural factor” effects, is, in fact, not helpful as it impedes serious proactive discussions. Alternatively, the unhelpful “explanatory devices,” may actually limit consideration of serious and substantive queries having to do with (a) the nature of vulnerability, (b) the role of context and its persistency of compromising conditions, (c) the salience of culture as adaptive traditions, (d) the nature of assets required for resiliency, and (e) the myriad consequences of inadequate supports. Given James et al. (1992) theorizing, the lack of supports, in fact, may suggest a large and underacknowledged impact of John Henryism and influences due to the nature of microsystem experiences of youth. Additionally needed are insights about the salience of long-term challenges that represent persistent conditions of high risk in contexts which assume the same supportive impact of the American Creed for everyday experiences of all citizens. The varieties of conceptual dilemma concerning the American Creed and the lack of equal access noted diminishes the production and availability of authentic information required concerning the specific character of resources actually needed and adequate for African American youth resiliency. Also affected and minimally understood is the impact of misinformation on youth performance. The situation impedes the design and delivery of adequate supports. Strangely, the shortcomings suggested are not represented in research strategies, overall project conceptualizations, or methodological techniques implemented. Nor are the challenges described evident and considered in questions asked by social scientists for determining their contributions to the broadly disseminated disparity findings that impact funding decisions. The dilemma is particularly salient for the developmentally relevant, conceptually unavoidable, and cognition-linked perceptions inferred by study participants themselves. Although the impact of accrued individual insights might vary according to the group's status,

all the factors are critical for youth, irrespective of race/ethnicity and SES. In addition and quite recently, social bias denial (e.g., given the election of the nation's first African American president) provides yet an additional source of social stress requiring coping and adaptations to alleviate psychological, physiological, and social impact. Failing to acknowledge social and racial stigma linked to group membership remains problematic, especially in the absence of critical supports. The situation contributes to traditional “culture blaming” social science research explanations (see Small, Harding, & Lamont, 2010). Additionally, another postelection perspective suggests that the United States became “postracial” virtually overnight—that is, the American Creed is similarly experienced by all. At the least, the widely and rapidly disseminated assumption of a postracial America may have served to exacerbate the previously noted conceptual challenges. The broadly aired, continuing, and social media–assisted conceptual strategy of framing the United States of America as a postracial nation attempts to reverse the U.S. identity as a raceconscious society wherein the American Creed and everyday race linked practices are at odds. The long-term program of research demonstrating race consciousness beginning in early child development was highlighted by Kenneth and Mamie Clark (1939, 1940) and Horowitz (1939). Research replications and extensions (e.g., Spencer, 1976, 1985; Spencer & Horowitz, 1973;), racial identity theorizing (e.g., Cross, 1991, 2003), and synthesis theoretical statements continued over the next several decades; see work by Spencer (1985, 1990a) and associates: Spencer and Dornbusch (1990); Spencer and Markstrom-Adams (1990); and Swanson, Cunningham, Youngblood, and Spencer, 2009. The focus has continued into the current period, although the research is plagued by problematic methodological strategies (see Brown & Bigler, 2005) and Kinzler and colleagues (Kinzler & Dautel, 2012; Kinzler & Shutts, 2008; Kinzler, Shutts, & Correll, 2010). There is no question that even very young children are unavoidably aware of differences in values about and treatment by race. Changes in cognitive growth and development include the learning of stereotypes and objective concepts, given opportunities for learning the alphabet, colors, gender role expectations (representing similarities and differences), number recognition, as well as the treatment, values, and labels associated with racial/ethnic group membership; concurrently, the learning of “color-blind” attitudes by White children (e.g., see Brown & Bigler, 2005) might, in fact, belie their exposure to models and the learned expectations of White privilege (which is infrequently acknowledged or assessed in developmental science). With progressive increases in cognitive maturation during middle childhood and continuing into adolescence, progressively sophisticated analyses are evident and unavoidable, particularly with regard to color-associated evaluative judgments and attitudes and beliefs concerning group membership (e.g., see Fegley, Spencer, Goss, Harpalani, & Charles, 2008; McGee & Spencer, 2013; Spencer, 1990b; Swanson, Cunningham, Youngblood, & Spencer, 2009; Way et al., 2008). At the same time, there continues to be a penchant to ignore the significance of racial consciousness findings, maturation-associated patterns, contributions of such patterns to stereotyping and stigma, and myriad historical events leading up to the perspective of the twenty-first-century postracial America as the context for human development more generally.

Perspective and Significance Lincoln's signing of the Emancipation Proclamation in 1863 was an important legal achievement. Nonetheless, regarding its everyday manifestation and influence on contemporary American life, the decree fails to provide the implied equity and fairness-valuing beliefs, attitudes, and practiced character traits assumed as the achieved outcome formally called the American Creed. Shortcomings include slight awareness of and concern about the psychological impact and challenges encountered over the life course on individual and grouplevel social experiences and policy-relevant consequences moving forward. The fact of the conundrum and its stability over time aids in explaining the high-risk contextual experiences had for some and a status of “invisible privileges” for others. For the most part, and as another contributing factor to this conundrum, the state of affairs is not acknowledged in formal social studies curriculum and disparate race-relevant academic experiences. Nor is the fact evident in the varied systems purported to afford myriad supports. In fact, social situations described as unequal conditions communicate a peculiar North American history. Generally understated is another critical context-relevant development pressure on youth: the historical fact and persistence of inequality influencing the conduct of science. The shortcomings impact the research questions posed, the character of the theoretical and conceptual framing of the work posited, and interpretations of findings (as well as the conceptual framing of the scientific enterprise itself) (see Fisher et al., 1993, 2002; Graham, 1992; Guthrie, 1976). All these problems significantly undermine potential supportive impact (or not) of social policy and opportunities for fair practices intended to shore up equalizing conditions for all youth and, for the most part, their support-intending families. Consistent with the African proverb cited, a highly critical element of the persistent American inequality includes the limited conceptual framing of human development research and theorizing for those considered “other” (see Harris, 1995). For example, the design of significant numbers of studies on American youth of African descent is frequently troubling. As an example of one major research flaw, methodological strategies too frequently compare low-income minorities with middle-income Whites. One result has been a conceptual pattern for minority status to be viewed as synonymous with lower income status, no matter families' actual social class status. The consistency of this pattern reveals the American Creed fallacy; in fact, housing restrictions, employment discrimination, loan procurement traditions by banks, and “redlining” (housing) practices continue to control the neighborhood opportunities, including quality-oflife concerns (e.g., healthy food options and services), available to working-class and middleincome families of color independent of their actual earning capacity and income bracket. Accordingly, the attendance of nonpoor minority children in schools characterized by percent of participation in free lunch programs is never disaggregated from school-level labels (i.e., level of participation in free lunch programming). Among other disadvantages, the practice contributes to the assumption of homogeneity among residents within the Black community, an assumption that ultimately fuels stereotyping and synonymous (low) expectations. As suggested, notwithstanding long-term critiques of comparative practices in research (i.e., along with other similar methodological oversights), the undifferentiating and stereotypical conceptual associations made about minority life—and particularly the development of African

American children—continue as routine and troubling traditions (e.g., see Fisher et al., 1993, 2002; S. Graham, 1992; Guthrie, 1976; Spencer, 1990a; Spencer, Brookins, & Allen, 1985). The conceptual orientation and patterned research tradition encourages a level of superficiality of knowledge garnered about the context of human development processes for varied individuals of color as well as many diverse generational immigrant groups. These underacknowledged practices compromise perspectives about the risks and challenges confronted by children of color and thus normative levels of high vulnerability. Additionally, the frequent and a priori “strengths- or asset-free” character of such research minimizes contributions to non–problem-focused questions about life experiences of minority children and adolescents. Our approach posits that all humans are unavoidably vulnerable (i.e., are confronted by challenges and risks) but have access to myriad supports and unique protective factors (see Spencer, 1995, 2006, 2008a; Spencer et al., 2006; Spencer, Swanson, & Harpalani, 2015). Thus, given the balance between accessible supports versus challenges and risks, appropriate ways of describing situations include statuses of high versus low vulnerability. Resilient individuals are those burdened by high risks and challenges but are able to take advantage of high (and often unique) supports. Privileged individuals may have challenges and risks but simultaneously benefit from superordinate levels of accessible and numerous supports and protective factors. The history of the social sciences has generally associated minority status with high vulnerability conceptualized narrowly as high risk level; Whites and favored minority-status (i.e., particular subgroups of Asian Americans) individuals have been associated with “normalcy.” The privileged American status of the latter groups is generally unacknowledged. The former situation is an added source of social risk and conceptual challenge for research and policy on African Americans; the latter circumstance is viewed as an additional component of an individual's knapsack of privilege, as described by Peggy McIntosh (1989) and other critical race theorists. Historical critiques of research traditions on minorities have been explicit although generally insignificant impacts on the conduct of social science (e.g., see Guthrie, 1976). As indicated, contemporary methodological and conceptual dilemmas linked with the framing and implementing of research on minorities of color continue to represent a “flat learning curve” (e.g., Fisher et al., 1993, 2002; S. Graham, 1992; Spencer, Brookins, & Allen, 1985; Spencer & Spencer, 2015; Swanson, Spencer, Harpalani, Noll, Seaton, & Ginzberg, 2003). In fact, the dilemma is clearly illustrated by a recent large-scale, positive youth development (PYD) study (see Lerner et al., 2011). Representing an impressive scope, the long-term, large, and much-needed undertaking included virtually 7,000 participants, living across 42 states, and represented at least eight repeated sequential points of data collection; in fact, given the sensitivities and sensibilities of the collaborators, the multiyear national data collection project included a variety of minority participants. However, the implementation of even this carefully conceptualized study, which suggested a critical consciousness concerning minority representation, still failed to include adequate numbers of minority youth to successfully accomplish within- and between-group analyses. As indicated by the Commentary on the recent publication of project outcomes (see Spencer &

Spencer, 2015) and particularly given the PYD emphasis, what could be conceived as a research implementation failure represented much more than a missed opportunity. If analyzed from an ecological perspective (see Bronfenbrenner, 1992), which acknowledges the contributions of the chronosystem and thus historical forces on the character of ecologies navigated by young people, missed opportunities were numerous. These missed opportunities were particularly salient for African American male youth, given that preexisting circumstances require close attention to understand contemporary everyday inequalities and long-term representational inequities in the conduct of science. In fact, across the life course, living as a Black male is fraught with myriad and unique challenges (e.g., see Barbarin, 2013a, 2013b; Barbarin et al., 2013; Boykin, 1986; Cunningham, Swanson, Spencer, & Dupree, 2003; Lewis, Butler, Bonner, & Joubert, 2010). The perspective described and methodological inadequacies noted (e.g., sampling flaws) are consistent, as well, with the Einstein epigraph. Accordingly, if vulnerability is defined as a universal human characteristic wherein everyone has both protective factors or assets and risk characteristics or challenges, having available assets or accessibility to resources matters. That is, protective factors must be present and prominent to offset the dilemma and extraordinarily high-risk situations. When it comes to racial group membership in the United States, and as described by Cooley (1902) over a century ago, the context character for one's evolving sense of self matters. In fact, Cooley noted, “The thing that moves us to pride or shame is not the mere mechanical reflection of ourselves, but an imputed sentiment, the imagined effect of this reflection upon another's mind” (p. 152). Although represented as varying realities consistent with the individual's stage of development, mechanisms associated with biologically determined maturation experienced in multiple layers of social context are unavoidably unique, influential, and heighten awareness of self-relevant perceptions. For example, the egocentrism of young children associated with normal maturation protects the self from troubling race/ethnicity associated feedback of the type described by Cooley (i.e., particularly during the sensorimotor and preoperational periods of cognitive development) (see Spencer, 1976, 1982a, 1982b). However, as children move into middle childhood and adolescence, changes in inferencemaking capacities occur. Appropriate, expected, and diminished cognitive egocentrism takes place. Progressive cognitive shifts change the character of manifested egocentrism (see Elkind, 1966, 1967). This, in turn, reverses the ego-protecting function of children's developmentally appropriate level of cognitive development (see Spencer, 1985, 1990a). Thus, when linked with Cooley's perspective, Einstein's quote demonstrates overlapping and synergistic links between maturation-based cognition, socio-emotional functioning, and unavoidable human biological processes of youth; children's intellectual (cognitive) processes are associated with their emotional life, which are collectively connected recursively to biology and thus are unavoidable aspects of human development and functioning. Given the context of development and recursive experiences had with others—and as suggested by the Cooley (1902) quote—identifiability with and/or membership in a particular racial group or ethnicity still matters in twenty-first-century America. Accordingly, when considered long-term from a chronological perspective (the chrono-level of the ecology), the history and treatment of minority groups must be factored into analyses for achieving authentic

insights and understanding of group members' contemporary life experiences. Thus, history provides insights about the problem of persistent inequality—as context—and individuals' daily experiences. Appreciating the character and expressions of human vulnerability and resiliency of North American children and adolescents of African descent must be framed from a historical standpoint, which provides interpretive framing for contemplating needed supports, policies, and enveloping practices. That said, and as an opportunity to situate and emphasize the fact, race/ethnicity (i.e., identifiability with group membership) represents a source of significant risk for some (thus, increasingly high vulnerability) and, at the same time, corresponds with unacknowledged privilege for others (thus, lowered vulnerability). The next major segment of the chapter highlights the concepts of vulnerability and resiliency, providing a necessary perspective inclusive of all humans. In particular, it provides broadened perspectives and interpretations of development for youth of color. The section includes a discussion of the residual problem of slavery and its enduring character, which determines the context for human development while also serving as an influence on implemented policies and practices (i.e., the “what” and “how” of policies and practices intended to support and assist). The historical perspective also considers the persistency of privilege as Whiteness and African Americans' treatment as property associated with protected legal rights (see Harris, 2002) and describes contemporary colorism themes as well. The chapter discusses an ecological systems human development theory (i.e., phenomenological variant of ecological systems theory, PVEST). As an identity-focused cultural ecological perspective, PVEST is a unique life-course human development theory. Its context and culturally inclusive character accommodates the experiences of diverse humans (i.e., both minority and majority group members) as each navigates unique and traditional tasks associated with specific periods of the life course. The distinctiveness of the dynamic and recursive systems perspective lies in its ability to represent, address, and explain the mechanisms of human development (i.e., both as processes and as outcomes) of minority as well as majority group individuals; thus, it explains development of humans (i.e., not only individuals of color). Accordingly, as an identityfocused cultural ecological perspective, PVEST encompasses the minority life-course experience and majority group meaning-making processes, thus affording a strategy for examining the vulnerability status and resiliency patterns of African American children and youth while, at the same time, acknowledging the role of the American Creed and historical contributions that reified “Whiteness.”

Human Vulnerability and Resiliency Before providing as background the role of history in the everyday contemporary lives of African American youth, a discussion of vulnerability and resiliency is critical. Consistent with all stages of the life course and as suggested elsewhere (Anthony, 1974; Spencer, 2006; Spencer, Harpalani, Cassidy, Jacobs, et al., 2006; Spencer & Swanson, 2013; Spencer et al., 2015), adolescence unfolds in contexts influenced by the particular historical moment, physical space, sociocultural traditions, and distinct interactions among them. As part of their efforts to achieve adulthood status, adolescents strive to attain progressively abstract academic and

career-linked competencies as well as socioemotional-relevant relational skills. Their efforts may occur in generally underacknowledged conditions of stress and risk as well as with broad arrays and levels of support. Nonetheless, expectations for youth stage-specific outcomes and adulthood-supporting competencies are broadly disseminated by the media, which also aid in their reinforcement and maintenance. Thus, the socially constructed and historically linked variations in levels of support are frequently overlooked or not recognized in media portrayals of the many within- and between-group gap findings. Group membership as social identifiers, including gender, race, immigration status, skin color, ethnicity, and socioeconomic position, influences the availability and accessibility of resiliency supports. The American Creed frequently prevents candid discussion of social differences in group-associated treatment. The absence of authentic ways of acknowledging the salience of identifiability and stigma is problematic. Accordingly, also missing are options to discuss the impact of the shortcomings and unavoidable implications—to name a few—such as the training of professionals (e.g., teachers, health professionals, criminal justice personnel), the conduct of research (e.g., the nature of the sampling technique), the design of programming (e.g., social and health delivery services), and the framing and ultimate implementation of policies. The exemplar categorical associations noted impact statuses of privilege and, at the same time given one's particular categorical position(s), support insufficiency. These associations are both alarming and informing. The possible impact of supports and challenges for groups of individuals have the potential to widen gaps by further privileging situations for some while simultaneously and potentially adding to the challenging conditions endured by others. The theoretical stance taken emphasizes the shared human experience of vulnerability. The level of vulnerability indicates the balance (or imbalance) between protective factors accessible versus risk level present; both vary, and thus human vulnerability shows significant plasticity. The fact of human plasticity reinforces the importance of accessible supports and the impact of risk conditions (i.e., both perceived and in fact). Accordingly, our view posits that as a life-course fact, all humans experience a variety of risks as well as secure benefits from protective factors of one sort or another as each individual navigates the world (see Spencer, 2006, 2008a; Spencer & Swanson, 2013; Spencer et al., 2015; Spencer & Tinsley, 2008). The framing suggests that human vulnerability does not reflect most traditional research perspectives but instead, consciously or unconsciously, situates privilege (including generally unacknowledged benefits) as the status of all citizens. The assumption implied is one of equal accessibility to supports, thus implicitly reifying the genuineness of the American Creed for all. More specifically, equity as fair competition and equal opportunities for advancements are seen as uniformly attainable. Given actual lack of credibility and legitimacy of this view for the everyday lives of particular citizens (i.e., especially those of color and/or highly impoverished individuals), the assumption demonstrates shortcomings. The inference of equal access promotes the conclusion that disproportionality or failure for some, and the patterned successes for others, are legitimate and “earned” achievements or group status outcomes. That is, some of a society's citizens, as success stories or assumed earners of superior performance, are portrayed as the norm. Individuals' disproportionate access to protective factors and supports (i.e., those enjoyed

without full acknowledgment of them) is seldom noted. Given the inequalities that characterize experiences by group membership, theorizing for those having experienced such sensitivities has been helpful.

An Inclusive Perspective: Phenomenological Variant Of Ecological Systems Theory Providing new conceptual strategies of salience to all humans, phenomenological variant of ecological systems theory (acronym pronounced: P-VEST) is a human development framework that acknowledges the critical role provided by individuals' own perceptions (Spencer, 1995, 2006, 2008a; Spencer, Harpalani, Cassidy, & Jacobs et al., 2006; Spencer, Swanson, Harpalani, 2015). It focuses on identity formation while considering structural factors, cultural influences, and individual perceptions of self, significant others, life experiences, and the environments in which one lives and copes; the processes unfold for individuals or groups while they navigate across myriad social and physical contexts. As such, the system's framework is conceptualized as an identity-focused cultural ecological perspective, which represents significant and unavoidable plasticity. Identity formation takes place across the life course and is especially relevant for adolescents, given their heightened self-consciousness during that period of development. PVEST combines its emphasis on individual perceptions with Bronfenbrenner's ecological systems theory (1979), thus linking context and perception. While Bronfenbrenner's model provides a means for describing the ways in which multiple levels of context can influence individual development, PVEST illustrates life-course development within and between contexts and group membership. In doing so, it emphasizes the individual's meaning-making processes that underlie identity development and responsive behavioral processes and outcomes (Spencer, 1995, 2006, 2008a); both are linked with the developmental tasks and maturational themes salient for the particular developmental period. As typical and atypical risks and challenges are confronted, the character and level of accessible supports play critical roles. Supports and challenges matter particularly for youth development. The theoretical perspective of PVEST helps situate work by McIntosh on privilege and the concept of John Henryism detailed earlier. McIntosh (1989) designated the frequent (but unarticulated) association of support with high status. Characterizing the advantages as having to do with “Whiteness,” she described the association of Whiteness with privilege and indicated an individual possessing or having access to (or not) a knapsack of privilege. McIntosh suggested that benefiting from a “knapsack of assets and privileges” makes a difference; appropriating greater success with less effort, developmental tasks are handled with greater ease. However, the lack of easy access and “assumed privileges” and greater exposure to attendant tension (i.e., given minority-group status and social stigma) is infrequently acknowledged or considered in programs of research or the evaluation of research. Importantly, patterns associated with persistently high-risk status and hurdles related to multiple contexts matter as well. Additionally, the dilemma associated with high risk status, in fact, multiplies the problem of patterned adversity, given that significant challenges are perceived as “ordinary” life-course experiences. The dilemma virtually explains the physiological adversity described by James et

al. (1992) as “John Henryism.” More troubling, assumptions concerning legitimate outcomes of meritocracy accompanied by assumed parallel individual-level investments of labor, hard work, and intrinsic motivation (believed expended) reinforce stigma; they compound the state of privilege as “norm level” and reaffirm assumptions concerning effort expended for particular—and, in fact—privileged individuals. For some, widening outcome gaps reinforce stereotypes of laziness, lack of motivation, and intellectual deficits assumed to represent native capacities for the particular undervalued referent group. In contrast to these views, a conceptual commitment to the universality of human vulnerability and its dual contributors as protective factors and risks allows for greater exploration of the “how” of outcomes and not just the “what” as a development-specific product. The suggested alternative conceptual frame connects risks as pitted against individuals' diverse strengths and protective factors. The PVEST framing allows for consideration and analysis of particular levels (and character) of supports and protection. The framework affords a nuanced analysis that makes explicit both culturally specific and shared protective factors, strengths, and supports (e.g., intergenerationally based resources, religious beliefs, social connectedness, spirituality, faith traditions, cultural socialization practices, social connectedness, etc.). Also critical and a conceptual benefit of the framework is that a group's high vulnerability due to significant social risks (and neglect) are acknowledged (e.g., reversed policy impact, low school resources, significant levels of stigma and stereotyping, inadequate health resources); the emphasis is different from that of traditional social sciences and specifically developmental stances, which ignore their salience (see Spencer & Tinsley, 2008). In fact, Spencer & Tinsley (2008) indicate that there are at least two strategies for interpreting standards upon which programs of research are implemented and findings are evaluated and interpreted. One considers primarily protective factors for certain youth and groups; in parallel fashion, strengths of other youth are viewed as basically inferior or invisible in regard to salience. The latter group of young people, in fact, are represented in the literature as possessing primarily and narrowly defined risk factors; their availability of protective factors is not acknowledged (e.g., both as familial and group traditions and shared strengths). Accordingly, this shortsighted perspective is associated with assumptions of meritocracy for some and presumptions that high-risk-status individuals are devoid of strengths and supports. The clear bonus of the PVEST conceptual strategy is its inclusiveness of experiences representing culturally diverse communities and making explicit how the traditions are protected or challenged by specific opportunities or burdens. Accordingly, these research values or perspectives contribute quite different insights into the policy, research, and practice (e.g., as pedagogy) traditions which are present at any historical point in time, impacting life course outcomes. The conceptual strategy and framing of PVEST not only explore and acknowledge significant imbalances in human life experiences but potentially introduce significant dissonance when one reflects on the conduct of social science research, theorizing, and everyday policy-relevant practice. Challenges to Youth Development Irrespective of supportive conditions, biological, cognitive, and affective changes still have

the potential to enhance youth vulnerability, thus contributing to findings that suggest typical or atypical patterns of outcomes. As described, this fact also suggests that human vulnerability includes the relationship between frequently unacknowledged risks and frequently ignored privileges as protective factors. Supports matter, given the obligation to confront the requirements of developmental tasks (see Havighurst, 1953) and the psychosocial needs to support them (see Swanson, 2010). There are few points in life beyond adolescence (as well as the neonatal period) when the invisible or unacknowledged fact and character of vulnerability have more dire implications. Erik Erikson (1950) described human development across the entire life course and emphasized the need to focus on ego identity development as the defining characteristic of adolescence (e.g., see Muuss, Porton, & Velder, 1996). Erikson (1950) explained that “psychosocial development proceeds by critical steps —‘critical’ being a characteristic of turning points, moments of decision between progress and regression, integration and retardation” (pp. 270–271) and proposed eight distinct stages of development. Police officers may perceive Black males' extended good-byes with friends or with group of youth as threats, no matter how they are intended. Being experienced as a threat is associated with Black males understanding of how they are perceived by the world around them, thus shaping turning points relevant in their development (Stevenson, 1997). For Erikson, each stage is characterized by a conflict between two opposing forces; the primary goal of development is the acquisition of ego identity. Stage 5, which occurs during adolescence, is defined by a conflict between identity and role confusion. Strategies by which adolescents resolve the identity conflict impacts their development through the final stages from early to late adulthood. With the onset of puberty, adolescents are “primarily concerned with what they appear to be in the eyes of others compared with what they feel they are, and with the question of how to connect the roles and skills cultivated earlier with the occupational prototypes of the day” (p. 260). Given the problem of unequal access to opportunities for minority youth (African Americans particularly), stage 5 experiences for African American youth might provide quite different identity resolution opportunities. Without question, an accrued sense of ego identity is accomplished when “inner sameness and continuity prepared in the past are matched by sameness and continuity of one's meaning for others” (Erikson, 1950, p. 261). The danger of stage 5 is role confusion, a state in which individuals temporarily overidentify with peers held in high regard and, in doing so, seem to lose their own sense of identity. As a defense against role confusion, adolescents tend to be “clannish, and cruel in their exclusion of all those who are ‘different,’ in skin color” (p. 262) and other arbitrary characteristics. The differences in accessible supports matter. At the same time, inordinately high levels of challenge and stress variation make the identity processes (i.e., given coping supports) quite different for youth even if they appear to share “similar contexts.” For example and relative to identification processes, the election of White presidents has not represented the same challenges and assumptions about “Whiteness” for European American youth as it has for people of color, given the nation's election of its first president of African descent (e.g., broad media-disseminated assumptions of the achievement of a postracial status hours following the election of Barack Obama). To further illuminate the complex processes experienced during Erikson's fifth stage, James

Marcia (1966, 1980) described and empirically validated four distinct modes in which adolescents react to the identity crisis. The four modes differ along two variables—whether the subjects have undergone a crisis and whether they have committed to a particular occupational, political, or religious ideology. Identity achieved individuals have experienced a crisis or exploration period in which they were forced to reevaluate their beliefs and have emerged from the crisis with a commitment to a particular identity. Exposure to myriad privileges would be expected to impact the process. Foreclosure individuals have also committed to particular ideologies or occupations but have not experienced the crisis period. These individuals tend to exhibit identities that mirror those of their parents or other socializing adults; these identifications are potentially dissonance producing for youth of color, given the extra stress and strain on parents and a potential John Henryism level of effort required for carrying out societal roles (see James et al., 1992). Thus, the greater challenges associated with minority status (including relational processes with parents) makes the identification processes of minority youth more nuanced (at best). McIntosh's (1989) knapsack analogy suggests that high-privileged parents may experience adult developmental tasks differently, given their access to unacknowledged privileges linked to Whiteness. Obviously such experiences are quite different from the support insufficiency situation described by James et al. Moratorium individuals are actively engaged in the adolescent crisis period; that is, they explore different paths and question belief systems. From moratorium experiences, individuals may reach a point of commitment and identity achievement or may disengage from the crisis without making a commitment, thus entering the next stage, identity diffusion. Of course, the extrarelational challenges associated with minority status indicate that the comfortableness of engaging primarily in experimentation may not be an option for youth of color. Identitydiffusion individuals are characterized by their lack of interest in committing to a professional, political, or religious path. For youth whose participation may be associated with membership in a stigmatized group or living under support insufficiency conditions, their options may be more limited; thus, diffusion may well be more constrained or show unexpected variations for youth of color. Not surprising, atypical trajectories may be observed more often, and researchers who perform uncritical analysis may infer simplistic conclusions that further promulgate stereotypes. According to Marcia's (1966, 1980) identity status framework, identity achievement is a dynamic process, which means that an identity-achieved individual can regress to a state of moratorium many times over the life course. When the environment changes and individuals are faced with dissonant value systems (e.g., questionable and dissonant feedback about the self and beliefs about what is possible), they may return to a state of moratorium. At any point, the moratorium individual has the potential to become either identity achieved or identity diffused. These shifting processes require psychic energy and supportive environments for productive resolution. Of the four modes of Marcia's framework, identity achievement is associated with the highest self-esteem and greatest resilience in the face of negative feedback. It is important to avoid value judgments when comparing foreclosure and moratorium. Each may be adaptive, depending on social context and individual differences. However, the level

of misinterpretation of the achievement gap fomented by Fordham and Ogbu's (1986) “acting White” characterization provides a case in point. Alternative interpretations have posited identity-focused but culturally and ecological-sensitive perspectives relevant to meaningmaking processes of youth. The more general historical events reportedly experienced immediately following the end of slavery suggest a perspective not consistent with the Fordham and Ogbu perspective (see Spencer, Cross, Harpalani, & Goss, 2003). As described by Cross (2003) and by Siddle-Walker (1996), historical documents as contextually relevant evidence suggest an overall valuing of education by Black educators and communities. Cited historical data and research reported by Cross and by Siddle-Walker reveal a general activist orientation by Black educators in support of the development of competence formation among youth. The observations are generally not included in the design of research for implementation, which emphasizes the role of teachers (i.e., attitudes, beliefs, social and emotional connectedness, and values) and may aid in explaining the findings of the atypical late-twentieth- and early-twenty-first-century achievement gap. Another strategy that links contextual factors with typical and atypical findings for Black youth is Boykin's (1986, 2000) analysis, which makes clear the potential impact of the context on tensions experienced in learning environments. His approach describes the source of and reason for heavy psychic energy misuse as the “triple quandary.” Boykin's characterization is consistent with PVEST (Spencer, 1995, 2006, 2008a); he reaffirms the role of cognitive mechanisms, coping processes, and identity pursuits. Boykin describes the three elements of the achievement triple quandary identity dilemma as the need of youth to balance identities; the three identities include a “mainstream White experience,” the minority [status] experience, and the specific (and frequently highly devalued) Black culture experience (see Boykin, 1986, 2000). Boykin's analysis, which is focused on identity and coping, suggests that the issue may not be youths' association of school-linked competencies with “acting White” identifications. From a PVEST perspective, the efforts of African American students indicate ego resiliency attempts to protect the self, given cultural values and the lack of contextual experiences that provide for a seamless inclusion, for example, in their school. According to Spencer and Harpalani (2008), Fordham and Ogbu (1986) were not the first to characterize the minority achievement gap as “acting White.” That perspective had been argued over a dozen years earlier by McArdle and Young (1970). Counter to the perspective emphasized by Ogbu and Fordham, Spencer and Harpalani suggest that “gap findings,” in fact, represent a virtual “youthful pushback” from values designed to demean, devalue, and disgrace. Youth of color, frequently in environments of stigma and stereotyping, push back from accepting school-linked preferences and values. More specifically, youth consider the oftdominating values as focused efforts to denigrate Black culture, values, and beliefs (Spencer & Harpalani, 2008). Spencer and Harpalani concluded that youth need to resolve the counteridentifications noted (i.e., consistent with the triple quandary analysis proffered by Boykin) while simultaneously coping with school-relevant competencies. In regard to the acting White perspective, it appears fair to speculate that at any given time, youth have a finite amount of psychic energy for responding to school-relevant developmental tasks. The emotional combating of self-relevant (mis)labeling by youth represents a significant

misuse of psychic energy. Efforts by African American youth to protect their “cultural heritage” and race-relevant “selves” is energy consuming. The process may leave less psychic energy available for satisfying or resolving the tasks at hand (i.e., an academic learning orientation, productive school record, and achievement outcomes). Further, and as implicated by Marcia's (1966) identity status perspective, identity processes for stigmatized group members require more adaptive processes (i.e., both reactive and stable coping) and, accordingly, would be expected to be significantly and behaviorally nuanced. And as a function of unavoidable plasticity, adaptive processes suggest many more opportunities for producing atypical outcomes relevant to ego identity. Commensurate with the reactive and stable adaptive coping processes required when framed by a PVEST perspective, achievement outcomes for Blacks are often “explained” by “event labeling” (e.g., acting White or stereotype threat). Event-specific labels fail to acknowledge the complexity of the process and the contributions of supportive cultural traditions (e.g., religious identity [Brittian & Spencer, 2012]), the designation and importance of “referent others” (e.g., an elderly, culturally affirming next-door neighbor) who represent important sources of cultural identification and allegiance (see McGee & Spencer, 2013). In addition to informing convictions regarding careers, religion, and politics as normative tasks toward productive citizenship, racial and ethnic minority youth must also adopt an identity congruent with their group membership that offers the foundation of group strengths in contrast to the stereotyped societal messages promulgated in the media. Various models exist for explaining and assessing the relationship of these identity processes on youth outcomes (Cross, 1991; Sellers, Smith, Shelton, Rowley, & Chavous, 1998). Although there are between-group influences and differences in how identity formation develops, nonetheless, there is consistency regarding the impact on viewing the world from a Eurocentric lens versus a secure racial and ethnic identity. The former situates youth for internalizing stereotypic messages, while the latter may contribute diverse perspectives without youth risking their own identity. Luthar and Lattendresse (2002), in contrast, describe maladaptive coping of privileged youngsters who lack adaptive coping practice opportunities. However, from a developmental science perspective, few studies examine privilege, “fear of differences,” and commensurate identity processes in the case of White, middle-income youth. Thus, the contributions of historical circumstances, particularly the institution of slavery, assist the current conceptual task by linking the history of North American slavery to contemporary and frequently highly challenging experiences of youth and the impacts of Whites' perceptions of their own privilege as it functions and represents the context for others' human development. Another interesting revelation is the infrequent acknowledgment of the significant and positive impact, historically, that the Black community had on its residents and their children (see Siddle-Walker, 1996). In contrast to the historical role of the Black community and Black schools as described by Cross (1991) and Siddle-Walker (1996), the lack of an ego-supporting environment for Black youth in contemporary school and neighborhood settings is described by Stevenson (1997) as resulting in youth feeling “missed, dissed and pissed.” That is, no matter the historical relationships of Black youth in their schools and neighborhoods (see Cross, 2003; Siddle-Walker, 1996), legitimate inferences and unavoidable meaning- making

processes of contemporary youth in hostile environments are important (e.g., see Chestang, 1970). Such cognition-, psychosocially-, and biology-dependent processes matter for oftreported findings of achievement gaps. The situation serves to compound school-associated relational and identity challenges for both students and families.

Historical Placement It is important to acknowledge that the 2015 publication of the third edition of the Developmental Psychopathology handbook also marks the 150-year anniversary of the passing of the Thirteenth Amendment to the Constitution, which marked the December 1865 abolishment of American slavery. The maintenance of the institution of slavery contributed to significant levels of dissonance. Historical Relevance of Slavery for Contemporary Cultural and Social Traditions The institution of slavery was viewed by Thomas Jefferson, one of the nation's Founding Fathers, as a necessary evil. The views held about African slaves and their economic benefit, particularly to the South, resulted in the establishment of the Confederacy and the waging of the Civil War. There is no doubt that slavery was seen as an economic benefit but also represented frequent concerns about potential social consequences of freed slaves, given the less-thanhuman labeling assigned to them. It is critical to acknowledge that the passing of the Thirteenth Amendment to the United States Constitution in 1865 was an enactment of policy. It did absolutely nothing to contain or change the broadly held assumptions, values, assessments, and beliefs concerning people of color as fellow humans residing as equal citizens of the nation. As suggested by the Chestang (1970) analysis, the inhumane labeling solidified the socialization task for parents of Black children to engage in supportive parenting and positive youth development efforts, albeit under multilayered hostile ecologies and practices regarding home, school, and neighborhood experiences. Situations and strategies to change thought processes concerning color-associated “difference” were neither acknowledged as fundamentally needed nor formally enacted. In sum, policies were fomented and filed as the law of the land, which—of course—was no small feat. However, continuing attitudes, beliefs, and feelings about the price paid for the freedom of “unlike others” were left virtually unaddressed for over 140 years. More specifically, during the recent half dozen plus years (and following the passage of the Thirteenth Amendment in 1865), several reconciliationrelevant resolutions were passed. In 2007, the State of Virginia Assembly was the first state to pass the House Joint Resolution (Number 728) acknowledging “with profound regret the involuntary servitude of Africans and the exploitation of Native Americans, and call for reconciliation among all Virginians.” On July 30, 2008, the United States House of Representatives passed a resolution apologizing for American slavery and called for reconciliation for the subsequent discriminatory laws. The United States Senate followed suit on June 18, 2009, also apologizing for the “fundamental injustice, cruelty, brutality and inhumanity of slavery.” However, assuming no real damage or long-term harm, the latter statement also noted that the resolution passed could not be used for requesting restitution. It is

important to concede that before the Virginia resolution in 2007, 142 years passed when absolutely no responsibility was taken for the nation's institution of slavery and treatment of Blacks during the interim. Further, one might infer that the general opinion continues to be that the sustained beliefs and traditions about people of color during the interim were not salient in the design of supports, or contemporary race relationships, requiring redress of any kind. Unlike the end to the institution of American slavery, at the end of apartheid in South Africa, truth and reconciliation public hearings provided opportunities for psychological processing of its long-term and heinous character as an unjust and brutal social and political system. The tribunals at Nuremberg at the end of World War II served a similar purpose. In fact, contemporary efforts continue to support the efforts of kin to obtain valuables confiscated from ancestors brutally killed in Europe. As noted, however, and for over 140 years following the end of slavery and passage of the Thirteenth Amendment, responsibility was never assigned nor guilt acknowledged moreover than political resolutions generated and filed. The amendment provided rights but without accountability for protecting individuals' rights against exclusionary practices. Accordingly, one perspective about the persistent adverse attitudes and identity construal of Black Americans is that unaddressed and residual attitudes, assumptions, and beliefs continue to contribution to contemporary contextual and psychological conditions. The unacknowledged circumstance results in the enjoyment of privileges for some and enduring threat as sources of significant everyday risk and challenge for others (i.e., leading to multilevel support insufficiency). The conundrum has implications for practices and traditions across the life course and at all levels of social ecologies, including implications for conducting research, implementing practice, and shaping policies. For the nation's youth, the lack of attention has added to the complexity of experiences evident both within and between groups' everyday developmental experiences (i.e., particularly youth of color). Identifiable resources and access to them matters; thus, protective factors such as Whiteness for Caucasians and skin tone differences for African Americans (as well as cultural socialization experiences needed for coping with persistent and nuanced discrimination) have implications for the character of treatment, risk level confronted, and accessible or experienced supports. Color and race still matter for African American youth. Historical Framing of Challenge and Opportunity from a Critical Race Perspective Critical race theory affords a late-twentieth-century and early-twenty-first-century lens for framing the design and interpretation of research on African American youth (i.e., particular patterns of findings for typical and atypical outcomes). For our purposes, it also provides a strategy for utilizing historical events for interpreting twenty-first-century findings about developmental outcomes for minority youth (i.e., both expected outcomes and more atypical ones). The use of historical sources supports the centrality of particular opportunities and challenges regularly confronted by youth of color. At the same time, the historical backdrop provided and critical race perspective, collectively considered, afford explanations for the particular types of opportunities and challenges confronted by Black youth. Additionally and of

relevance to resiliency findings, critical race considerations provide useful ways of thinking about salient constructs for consideration in programs of research but too frequently remain conspicuously absent. The conceptual deficiency disguises and complicates authentic comparisons between the daily reality of minority group youth versus those of nonminority youth, who customarily enjoy or benefit from what has been referred to as “Whiteness as property” (Harris, 1995). In this case, the term property suggests experiences including opportunities that are governed by social rules; relatedly, the North American brand of slavery had unusually horrific social rules. In fact, and consistent with this perspective, D. B. Davis (1966) described the continuity of slavery as a concept as well as a social and legal category. Operational Whiteness traditions separated Blacks and Whites relative to slavery. However, Whiteness itself became associated with sets of social rules and traditions in the mode of property, as described by Harris (1995); from this lens, Blacks became objects of property (i.e., Whiteness). Davis suggested that slavery in the modern Americas was remarkably dissimilar from other forms of slavery; its uniqueness had to do with its importance as an economic commodity in the plantation economies. Thus, that importance was a key reason for its distinctness. Perhaps its uniqueness and continuation into the twenty-first century is due to its continuing economic impact. More specifically, failing to confront the view of Whiteness as property suggests that the dire outcomes for Blacks functions as an economic system (i.e., job creation, which sustains the contemporary form of an earlier tradition). Careers can be devoted to “fixing others” or the design and implementation of policies and practices misconceived and inadequate as supports for guaranteeing opportunities for self-determination among Blacks. The social rules that were altered with Lincoln's signing of the Emancipation Proclamation represented a change in a legal tradition. However, it might be argued that the social rules governing everyday access to power and the control of power over nonmaterial resources (e.g., access and opportunity) did not change; the persistency may explain the current design of research, the queries posed, and the interpretation of everyday experiences had and required of youth of Black families and communities. Further, given the lack of change regarding perceptions of and representations of Whiteness as property, James et al.'s (1992) epidemiological relevant observation of “John Henryism” is not unexpected. That is, the Whiteness-as-property social dilemma, in fact, would make the resources and supports actually required for offsetting the level of effort expended virtually unobtainable in accomplishing “ordinary developmental tasks” as described by Havighurst (1953). Considered collectively, the critical race perspective (e.g., see Harris, 1995) evoked and the continuing residual problem of (social and psychological) slavery as described by D. B. Davis (1966) afford new insights concerning human vulnerability and resiliency when applied to the experiences of contemporary African American youth. As suggested and from a developmental science perspective, there continues to be a tradition to ignore the normative knapsack of White privilege (McIntosh, 1989) and general salience of the Whiteness-as-property thesis (Harris, 1995), a tradition that communicates a litany of intergenerational privileges. The socially constructed supports and rules—unacknowledged but operational—automatically provide White youth and families with unacknowledged social and personal supports; they virtually guarantee positive “typical” outcomes; further, their utility

operates to decrease the obvious sources of high vulnerability (i.e., especially economic sensitive indicators). From a critical race perspective, a legally supported Whiteness-asproperty tradition has dire consequences for minority resiliency, self-determination, and beliefs about the American Creed. This “silent” although operational tradition contributes to persistently low human vulnerability for nonminority citizens independent of social class. As cited by Harris (1995) and specifically analyzed by Roediger (1991), there are wages (such as having consistent benefits and unacknowledged assets) associated with Whiteness. Importantly, and given the obvious consequences and benefits of not acknowledging the benefits of Whiteness, bringing attention to the different contributors to and consequences of it for human vulnerability matters. Roediger's perspective also assists in explaining the differences among poor White, Black, and Brown individuals; that is, poor Whites may persist no matter their situation since being White still “pays wages.” In similar terms as Roediger and as McIntosh, Harris suggests that economic challenges may be experienced differently by poor Whites even during conditions of economic downturns. Harris noted that “Whiteness retains its value as a ‘consolation prize’: it does not mean that all whites will win, but simply that they will not lose, if losing is defined as being on the bottom of the social and economic hierarchy—the position to which Blacks have been consigned” (p. 286). Long-term and consistent stereotypes are associated with the position and have important significance for performance (see Aronson, Lustina, Good, Keough, Steele, & Brown, 1999; Aronson, Fried, & Good, 2002). No matter the effort expended, unless having and opting for the “opportunity” of “passing,” the Black vulnerability level is never made less severe by such reassurances as described by Harris—that is, of never being at the bottom. That fact, in and of itself—and independent of one's point in the life course—promotes a very different level of stress and risk and approach to everyday challenge. Maturation-based social cognition makes the awareness of dominant race-as-caste messages evident beginning in the early years. Such messages have been recorded as young children's race awareness, racial attitudes, and racial preference literatures (e.g., see Spencer, 1985; Spencer & Dornbusch, 1990; Spencer & Markstom-Adams, 1990). Identity-relevant accommodations by middle childhood and adaptations at adolescence and young adulthood have also been reported (e.g., see Harpalani & Spencer, 2009; Harpalani et al., 2013; McGee & Spencer, 2013; Swanson & Spencer, 1999; Swanson, Spencer, Dell'Angelo, Harpalani, & Spencer, 2002; Swanson, Spencer, & Petersen, 1998). The cognition-based awareness of African American youth of Whiteness-as-property themes require that they cope with extrachallenging—albeit generally “invisible”—identity-relevant developmental tasks. Although frequently understated in the literature, youth and families of color require more complex cognitive and emotional adaptations and context-relevant accommodations and supports (e.g., home, school) (e.g., see Brittian & Spencer 2012; Cross, 1991, 2003; McGee & Spencer, 2013; McGee, Hall, & Spencer, 2015; O'Connor, Hill, & Robinson, 2009; Spencer, 1990b). The knowledge base and resource needs for parents who must serve as effective socializing agents are different from those of parents who enjoy the social privilege of Whiteness as property. As described by Harris (1995), a fundamental and core valued aspect of privilege is “its exclusivity; but exclusivity is predicated not on any intrinsic characteristic, but on the

existence of the symbolic Other” (p. 290). Harris explained that this symbolic character also communicates an “illusion of unity” among Whites. Identity accommodations are not required by White youth as adolescents (i.e., coping with stereotypes about group membership); that is, in general, White youth can pursue traditional developmental tasks of salience used for assessing stage-specific competencies and general resiliency outcomes (see Havighurst, 1953). The experiential differences between Whites and Blacks—based solely on skin color—matter for coping as well as adaptations required, including identity formation processes, which have implications for level of vulnerability. Consistent with the opening quote by Einstein, at the least, the findings of too many studies about minority children and families create highly troubling literatures. More specifically, as contributed to by “experts” in the disciplines and fields of sociology, psychology, education, public health, social welfare, criminality, and criminal justice, a tax dollar–funded “research industry” is frequently overly simplistic and represents decontextualized approaches. Considered collectively, the shortcomings create or contribute to a virtual “culture of poverty” literature. It provides problematic interpretations of developmental processes and outcomes for children of color as well as entire communities. The lack of understanding into developmental mechanisms and their linkages with long-term context considerations and historical traditions result in the dilemma implied by the Einstein quote, which has obvious implications for competency assessments and stereotypes about them (see Aronson et al., 2002; Steele, 1997). In fact, as suggested from a critical race perspective, such conceptual shortcomings reify and illustrate “the valorization of Whiteness as treasured property in a society structured on racial caste” (Harris, 1995, p. 277). One may infer that not knowing about or failing to acknowledge the character and heterogeneity linked to the unique racial caste system of diverse groups and their children's everyday experiences introduces a pattern of conceptual shortsightedness, at best. The situation influences ultimate questions posed that impact the nature and delivery of social policy, everyday cultural practices delivered in schools and similar settings (e.g., see Lee, Spencer, & Harpalani, 2003), and while navigating neighborhoods “protected” by individuals purporting to “protect and serve” but in fact frequently suggest historically different interactions with particular socializing agents (see Cross, 2003). Considered together, the implicit messages of the two quotes are consistent with the culture of poverty critique and analysis described by Small et al. (2010). The depression-like economic conditions of the last several years have impacted the United States and the globe. Accordingly, given the extreme competition for limited resources, the unacknowledged problem of unequal access serves to make worse the problem of inequality, thus increasing high levels of human vulnerability. That is, people of color who lack the property commodity benefits of “Whiteness,” as posited by critical race theorists, experience the economic situation differently from their privileged counterparts. For African Americans, in fact, it may be less clear whether further adverse economic impact is due to continuing biases that contribute to greater inequality or whether the “national recovery” lacks levels of support actually needed for communities of color (and its child development outcomes) to experience actual economic supports for offsetting resource or asset insufficiency. These

themes potentially impact and compromise youth resiliency, levels of human vulnerability, and interpretation of both. But they are not alone. In fact, particularly given the historical framing provided, the role and impact of race, ethnicity, skin color, identifiability, and ego processes are critical constructs. They represent a level of atypicality and conceptual uncomfortableness in the traditional approaches taken to social science research involving youth of color and particularly African Americans. The opening quotes suggest particular coping needs required by individuals and communities, given unavoidable inferred messages about group membership, particularly for minorities of African descent. The conceptual strategy is necessary both for establishing and maintaining a healthy “sense of self” and for determining and supporting—as effective policies and practices —positive adaptations to typical and atypical human development needs and issues. Bronfenbrenner's (1992) framework and the historical framing provided explains the problem of unnecessary coping requirements as a function of status difference associated with skin color variations. The situation indicates a normative state of lack of fit or unacknowledged privileges experienced as children prepare for, navigate across, and interact with determinative environments. The character of the interactions has consequences for challenges and/or opportunity. The research questions pursued and methodologies considered in the conduct of twenty-first-century programs of research, evaluation, practices, and subsequent policies must represent youth placement and subsequent experiences. The next two sections highlight and review some of the research on youth of color, given the themes noted, and present for analysis methodological issues of importance for research that aims to be salient for the developmental tasks, and interpretation of experiences and outcomes for African Americans. By integrating and framing the historical and privilege-acknowledging perspectives presented, the final section concludes with a discussion of methodological concerns necessary to frame future research and theorizing about African American children.

Salient Perspectives Required to Appreciate Challenges to Resiliency Given universal developmental needs and tasks, one quandary created by ignoring historical circumstances and treatment is not acknowledging that everyone has sets of developmental tasks to master. These tasks are associated with the individual's stage of development (Havighurst,1953) and are influenced by the chrono-level of the ecology; accordingly, the character of the social challenges faced by a society at a particular psychological moment in history has implications for outcomes achieved (e.g., the recent and—virtually—globally experienced economic anguish). Although citizens overall confront economic challenges, the nature of the experience for some may be decidedly different and thus exacerbating of risk level in many cases. Moreover, the fact of differential experience by race, ethnicity, and class (and interactions among factors including gender considerations for Hispanics and African Americans) is infrequently acknowledged and factored into the conduct of science and its interpretation and utilization as policy and practice in the design of culturally authentic supports. These are, in contrast, considered as control variables in studies that negate the

breadth of challenges experienced as a function of the interaction among the noted attributes. While appropriate as control variables in research designs exploring outcomes believed related to these factors, nonetheless, there remains less attention to the experiences that contribute to specified outcomes. The introduction suggested that different historical relationships between groups and untoward experiences borne by particular individuals have unexpected (and typical) effects on the character of the life course experienced. A basic feature such as the maturation status of cognitive and social development may, in fact, represent a significant dynamic and serve as a protective factor (e.g., the cognitive egocentrism of preschoolers) (see Elkind, 1966; Spencer 1982a, 1982b, 1985, 2008b; Spencer & Markstrom-Adams, 1990); however, during adolescence, maturational status of the noted domains may exacerbate relational experiences and may impact well-being of youth (see Elkind, 1967; Spencer, 2006, 2008b; Spencer, Harpalani, Cassidy, et al., 2006). That is, on one hand, young children's developmentally appropriate self-centered thought processes limit the internalization of risk and threat due to negative group membership stigma and general stereotyping. However, the interpretations and inferences made are more complex during middle childhood (e.g., Spencer, Cole, Dupree, Glymph, & Pierre, 1993) and adolescence (e.g., see Fegley et al., 2008; Spencer, 2006, 2008b; Spencer, Harpalani, Cassidy, et al., 2006; Spencer, Dupree, Cunningham, Harpalani, & MunozMilleret al., 2003; Stevenson, 1997). As noted in several papers, socialization strategies, beliefs, and assumptions matter as a protective factor as well (Spencer, 1982b, 1985). These themes, occurring at all levels of the ecology, have special salience across the life course (even within a lower-income caste [e.g., see MacLeod, 1995]) for African Americans and other frequently stereotyped individuals of color. We suggest that on a daily basis and in light of the historical and long-term conditions described, persistently occurring and uniquely patterned challenges characterize the everyday experiences of America's communities of color and should be included as part of the context for everyday functioning (see Szalacha et al., 2003). The challenges need not be necessarily framed as deficits and deficiencies but acknowledged as part of the context of everyday life. At the same time, for nonminority group members, representative constructs acknowledging privilege also requires recognition and inclusion in the conduct of research, particularly when between group comparisons are part of the methodological strategy. The dilemma of “invisibility” of contextual differences between groups matters. There are consequences for the conduct of and interpretation of findings; moreover, the research enterprise omits the recognition that the psychological and physical contextual differences have profound implications for individuals' basic meaning making, coping processes, and level of vulnerability; and, of course, there are varied experiences among ethnicities (see Szalacha et al., 2003). For example, the honorary White status accorded some Asian American ethnicities is generally not protested but is accepted as an underacknowledged protective factor (myriad privileges); of course, the situation introduces other types of within- and between-group tensions (e.g., see Diemer, Kauffman, Koenig, Trahan, & Hsieh et al., 2006). Consistent with global colonial historical traditions and practices, contemporary citizenship claims as well as immigration apprehensions—particularly in regard to individuals of color—introduce group-

level sources of stress, entitlement beliefs, and protective factors that remain ambiguous, anonymous, and, more generally—with a few exceptions—virtually absent from rigorous scholarly efforts in the social sciences. In fact, perhaps reflecting the nearly century and a half of silenced and unremorseful practices following the signing of the Emancipation Proclamation and the passing of the Thirteenth Amendment to the Constitution, traditions and beliefs became solidified (as well as “silenced” from consideration). Accordingly, robust behavioral and social science careers may represent successful constructions due too frequently to practices designed for reaffirming a priori assumptions of group inferiority (i.e., a pattern of scholarship that affords a virtual colonialist representation of others' lives). The possibility of intergenerationally transmitted attitudes and assumptions is important, given an articulated social science “objectivity goal” and training criteria that assume values-free strategies. However, practiced objectivity is less possible and requires more coping, given unacknowledged and unchallenged views about others' identities and placement within American culture, views that include long-term practices and particularly “colored and coded” definitions of humanity. This situation is exemplified and generally viewed as “typical” as structured approaches to topics such as student discipline problems in school contexts: that is, where punishment is viewed as a way to restore justice and deter future offenses (Wenzel, Okimoto, Feather, & Platow, 2008).The historically retributive approach strives to maintain objectivity and consistency but is unsuccessful in deterring undesired behavior. More relevant, however, to the issue of vulnerability is that the approach holds youth accountable for responding to socially constructed and underacknowledged contextual conditions. The character of outcomes is framed as implying incompetence and undesirability when, in fact, youth efforts and supports could never be an even match or represent a formula for success. Restorative practices are frequently used in educational contexts and are interpreted and examined as an alternative to systems of retributive justice (Amstutz & Mullet, 2005). As a prevention and intervention approach, the practices are based on the principles of restorative justice that involve accountability for all members of a relational community (Morrison & Vaandering, 2012; Parsons, 2005; Swain & Noblit, 2011). Schools' codes of conduct often require that students who break rules be subjected to punishments that often have little to do with their inappropriate behaviors or with those whom they have harmed by their actions (Kohn, 2006). Restorative principles are based on the premise that harming others damages interpersonal and communal relationships and that justice is best achieved by seeking to repair relationships through restitution, resolution, and reconciliation (Braithwaite, 2002; Morrison & Vaandering, 2012). Instead of delegating the resolution of rule violations to an impartial third party (i.e., a court, judge or jury, or in the case of a school, a dean or an assistant principal), offenders, victims, and the communities that have been affected (i.e., friends of the offender and victim, and community members) come together in an attempt to understand what happened, how it has affected everyone, and what can be done to ensure a supportive outcome (Braithwaite, 2002; McCluskey, Stead, Kane, Riddell, & Weedon, 2008; Morrison & Vaandering, 2012; Popa, 2012). Instead of expelling a student from the community (or suspending from school), the community seeks to reintegrate the student in a manner that also

holds members of the community accountable for supporting the student. This strategy is more consistent with the perspective proffered; atypically, the approach requires adults to examine conjectures regarding their personal assumptions and systemic practices to which youth utilize specific coping efforts to respond. Differences in the character of coping efforts needed as adaptations to everyday life course developmental tasks required for achieving competence and sustaining health and well-being for members of the human species are significant contributors to vulnerability level. Acknowledgment of the differences may serve as a national embarrassment and thus source of conceptual resistance for some individuals; however, at the same time, it represents an everpresent inconvenient hurdle for others. The fact is especially cogent for America's diverse citizenry, given the nation's reluctance to acknowledge the continuing salience of slavery, Jim Crow practices, and skin color salience in twenty-first-century life despite how long it has been since Lincoln signed the Emancipation Proclamation. As indicated, we suggest that the dissimilarity of experience and actual effort required has to do primarily with epidermal melanin level variation (skin color hue). Einstein's view of “human genius status” as the norm, in fact, is an enacted value of relevance primarily ascribed to Caucasian Americans. Although generally the bias is inadequately acknowledged in the social sciences, more often than not, it represents a consistent theoretical dilemma and frequent conceptual orientation of research efforts. The fact contributes to the character of social policy, educational pedagogy, and a “backpack of life course experiences” wherein Whiteness functions differently for groups and their members (i.e., as a burden, given its absence as representative sets of unacknowledged privileges is the dilemma); how to represent the problem of Whiteness as a knapsack of privilege as inequality in rigorous research continues to be the unaddressed task. Highlighting its determinative influence contributing to structural conditions, Whiteness status conveys high status and support for some citizens; and, at the same time, for those of a darker hue, the unacknowledged power of Whiteness solidifies nothing less than an institutionalizing of risk and challenge. Across the life course both for children and those intending to protect, support, socialize and train, an individual's particular Whiteness status makes a difference. It serves either to contribute to psychological consonance (i.e., affording a best-fit status between individual and context experiences) or facilitates the experience of dissonance, which requires extra consciousness, responsive energy, and time for obtaining resiliency. It also supports the achievement of positive youth outcomes in the face of significant and persistent challenge. The typical research approach is to assume that social address variables (e.g., social class and race categorization) adequately differentiate the complex experiences suggested. Accordingly, as conceptualized by Anthony (1987) and other scholars, vulnerability is a shared human condition. It communicates a level of balance or imbalance between individuals' degree of risk versus the level of protective factors either present or accessible. Accordingly, all humans represent a level of vulnerability, although the character and level of risks and protective factors may vary by type and level of significance. In fact and as suggested, a characteristic such as Whiteness may function as a source of privilege and support for some individuals (e.g., Caucasian Americans) and as a source of significant risk and challenge for others. However, as suggested, in the case of within-species melanin variation and societal

experiences, the fact of difference as daily practices experienced represents a set of atypical and unique events that parallel those alluded to in the Einstein quote. Further, particularly for the North American experience, the fact has always been responded to at the national level as a dilemma of everyday traditions and unexpected social policy outcomes, but never one of social justice. As historically evidenced, the failure as a nation to acknowledge and to accept responsibility for long-standing, intergenerationally impactful forces—both personal and group —at the individual and collective levels, chronic injury has consequences. Supports can come from sources infrequently acknowledged as constructs in research (e.g., see Brittian & Spencer, 2013; Dupree, Spencer, & Spencer, 2015; McGee et al., 2015; McGee & Spencer, 2013; Spencer & Swanson, 2015). Of course, like the failure of individuals to acknowledge the fact of the Holocaust or that the nation's first Black president is an American citizen— albeit with the name Barack Obama—as both a historic and contemporary fact, there have always been pockets of individuals comfortable with distorting history as a way of accepting dehumanizing practices for overall economic and individual-level gains. Everyday behavior, collective practices, and stigma-linked determinative policies based on flawed perspectives concerning individual responsibility continue to enjoy contemporary manifestations in research studies as constructs assessed and questions posed. The predicament described communicates and demonstrates the continuing advantages enjoyed by non-Brown individuals (i.e., described by Roediger as the “wages of Whiteness” 1991). The conundrum significantly adds to the persistent situation of unacknowledged privilege and, accordingly, a significant source of psychological support for those of a lighter hue, which is evident both within-group and between-group experiences and social practices. Specifically, curriculum based social studies lessons are taught from the perspective of “the victor”; as such, privileged students occupy a position of never having to explain or having dissonance about: (a) a group's enslavement (and often mortal destruction during the Middle Voyage) and the implications for perceptions about “the self” and one's community; (b) individual and group-level inhumane treatment, including the separation of families, raping of women, rulings against indigenous language use, and proscriptions concerning the teaching of literacy skills; (c) the creation and use of policies and practices determined by the dominant majority solely on the basis of either being of African descent or having one drop of Black blood; and, additionally, (d) as formally or informally communicated, White status situating a privileged self to never having to concede personal culpability in the heinous enslavement of a group, destruction of family systems of identification and support, murder, and—collectively as a nation—never having to confront or bear individual-level responsibility for the heinous institution of slavery itself. To reiterate, in many ways, the “free pass” set of conditions described and framed in American textbooks as historical representations of the nation's past represents what Roediger refers to the unquestioned “wages of Whiteness.” Considered overall, these themes are not included in the formulation of publicly and privately funded social science research. An excellent illustration is provided in the recent “thriving literature.”

New Concepts and “Traditional” Assumptions: Thriving and Development Thriving originally appeared in the literature as a set of positive “vital signs” (Benson, 1990). Subsequent research has aimed at identifying these signs and developing measures as indicators of thriving in adolescence, most notably the 5C's (Competence, Confidence, Connection, Character, and Caring). All of this work attempts to draw a distinction between thriving and happiness (Scales, Benson, & Roehlkepartain, 2011). However, for the most part, it does not acknowledge the themes concerning Whiteness status and privilege nor minority group membership—given the unique histories of each—in the construct's conceptualization. In fact, some scholars conceptualize thriving as the high point on a curve overtaking resilience, surviving, and succumbing (Carver, 1998; O'Leary & Ickovics, 1995). These writers connect adversity to the positive benefits that can come after a negative experience. Of course, the long-term and unacknowledged impact of slavery and unresolved between-group status issues do not represent the level of adversity considered. The work distinguishes between threats and challenges with the former producing harm and the latter often leading to improvements. Under this model, resilience represents a return to the status quo following the particular experience whereas thriving refers to the improved state that follows. These scholars address specific forms of thriving, including desensitization to the stressor, enhanced recovery potential from the situation, and heightened level of functioning. Although infrequently addressed under the rubric of unequal accessibility to supports that sustain recovery, as a recent construct, the perspective assumes an equally experienced American Creed and equity in individuals' access to and exposure to resources and supports. The ability to measure these abstract changes, however, is challenging, although some objective developments can be assessed. Concrete benefits, more easily measured, can include improved or newly acquired skills, greater confidence, and strengthened personal relations (Carver, 1998); all imply accessible assets and supports. Research has shown that the higher the number of positive developmental factors that a young person is exposed to, the more likely he or she will be to also report thriving outcomes (Scales et al., 2000). Generally and as suggested earlier, thriving may be applied to all youth, although accessibility is decidedly different by race, ethnicity, historical circumstance, and unacknowledged privilege. Lerner, Lerner, von Eye, Bowers, & Lewin-Bizan (2011) discuss the adaptive developmental regulations and bidirectionality of the individual and context that, if fostered properly, can contribute to an adolescent's thriving, something the authors define as manifesting “healthy, positive developmental changes” (p. 1107). What is not considered and included in the various analyses are the historical and contemporary experiences linked with group membership. In fact, the literature on thriving has been framed as the successor to the at-risk and pathologically focused studies of the past (Scales et al., 2011) with renewed attention on personal strengths and resiliency. However, the challenges and continuing risk factors are underexamined by race, ethnicity, and national origin. Historical contributions to the stability of inequality and sustainability of privileging status for Whites are never considered.

Those affiliated with the Search Institute link thriving to “sparks,” which they characterize as passions or interests that provide direction to one's life (Benson & Scales, 2009; Scales et al., 2011); however, in general, it is not mentioned that “sparks” may be conceptualized as opportunities that are unequally accessible. Fifteen indicators of thriving have been identified that assess qualities such as school success, social life, and emotional outlook. This work expanded on previous work that was connected to status-related indicators (Scales, Benson, Leffert, & Blyth, 2000; Theokas et al., 2005). Overall, these perspectives seem to indicate the importance of moving forward (Carver, 1998) through time as a critical factor in conceptions of thriving. Much of the literature suggests that thriving is less a state of being or trait but rather a process of change (Theokas et al., 2005) reflecting an active individual functioning across time and place who is able to exert flexibility and adapt to changing contexts. One difficulty, of course, is the difference between conceptualizing thriving as a process but measuring it at a moment in time (Benson & Scales, 2009). As suggested, unaddressed are the traditions and behavioral residuals of a nation with significant unresolved issues having to do with a history of slavery and a continuing contemporary condition of structured inequality and variously masked privilege associated with skin color. Growth is also implied in this framework, which poses a conceptual challenge in that it remains uncertain to what extent thriving can be distinguished from other forms of development (Carver, 1998). Carver (1998) concluded that thriving represents an “extreme” form of growth that operates similarly to other forms of growth, such as physical exercise, with the exception that the growth that drives from threats is unexpected. As presented in the literature, individual variables linked to thriving include those coming from both personal and social psychology. Specifically, this research is related to whether an individual can actually thrive in the face of any given adversity; however, as suggested by the chapter epigraphs, the conceptual orientation does not consider the persistency of beliefs, attitudes, and long-term unequal conditions that paint a very different picture of adversity as experienced by youth of color and African American youth (and males) specifically. Individual factors can include optimism/pessimism, hope/hopelessness, and self-mastery and efficacy. Contextual factors can include the availability/lack of social supports and the extent to which an environment encourages self-determination (Carver, 1998). Criticism of the thriving literature has focused on the inability for many scholars to come to an agreement about what constitutes a developmental asset and what contributes to and indicates thriving. Central to this dilemma has been the context and domain-specific nature of thriving (Benson & Scales, 2009). Without question, a conundrum is why investigators lack interest in collaborating in order to obtain the perspectives of individuals themselves. Of course, whether thriving should be treated globally inevitably depends on the nature of the indicators utilized; some (i.e., assuming leadership roles) clearly are culturally specific. Some have challenged the limitations of this literature on thriving and instead have sought to incorporate a racial, gender, and socioeconomic component into their work (Blankenship, 1998). Responses to this criticism have emphasized the importance of how many developmental strengths (i.e., sparks,

relational opportunities, and empowerment) youth are exposed to and how these are far better predictors of thriving outcomes than gender, race, and socioeconomic status (Scales et al., 2011). Scales and colleagues (2011) showed that adolescents who have sparks and supportive individuals in their lives are more likely to have other commitments and values that connect them to their communities. From a national sample of 1,817 (56% White, 17% Black, 17% Hispanic, 0.5% Asian) 15-year-olds, they found this was the case across demographic variables. However and unfortunately, the final study (Spencer & Spencer, 2015) could not explore the questions either between or within groups because of the inadequate number of minority participants. Similarly, self processes have been studied for at least 75 years, there are few conceptual and methodological constructs more misconstrued in the case of African American youth (for reviews, see Hare, 1985; Spencer, 1985, 2008a; Spencer, Harpalani, Cassidy, et al., 2006; Spencer & Horowitz, 1973; Spencer & Markstrom-Adams, 1990).

Black Identity and Self-Esteem Research into the association between racial/ethnic identity and self-esteem has tended to produce mixed results (Rowley, Sellers, Chavous, & Smith, 1998). Variations in ethnicity, geography, and age contributed to differences in these findings. Within the realm of Black identity, results have been more consistent with racial identity affirmation positively predicting levels of self-esteem in African American adolescents and college students (Resnicow, Soler, Braithwaite, Selassie, & Smith, 1999; Rowley et al., 1998; Toomey & Umana-Taylor, 2012). This has been particularly true in recent years as models have become increasingly more nuanced in assessing the degree and quality of racial identity (Rowley et al., 1998). In the first decades following school integration, research on self-esteem and race showed that in contrast to some widely held conceptions, Black children generally displayed higher levels of self-esteem than White children and that they did not suffer from the presumed self-esteem deficit (Hare, 1985; Simmons, Brown, Bush, & Blyth, 1978). Observations obtained by Simmons et al. (1978) drew from the results of multiple-choice tests administered to 798 students in the sixth and seventh grade across 18 schools in a large midwestern city. Subsequent work continues to support these findings (M. Hughes & Demo, 1989; Porter & Washington, 1989). Absent from this research, however, was how racial identity might predict or influence self-esteem levels in African American adolescents. Toomey and Umana-Taylor (2012) conceptualized two different ways in which racial/ethnic identity is linked to self-esteem. They drew on a risk and resilience framework in which racial identity can either promote self-esteem via pride in one's self/collective or protect by minimizing the negative effects of stressors via racial discrimination (Rutter, 1987). Phinney, Cantu, and Kurtz (1997) conducted a survey of 669 American high school students in order to assess the predictive functions of ethnic identity, defined as “a broad construct including sense of belonging, positive attitudes, commitment, and involvement with one's own group” (p. 170), on minority self-esteem. A series of regression analyses found that neither gender nor socioeconomic status were a strong predicator of self-esteem in African American youth but that ethnic identity was a significant predictor of self-esteem across racial groups.

Within the “protective” model, Yasui, Dorham, and Dishion (2004) found that self-esteem was associated with adolescent adjustment for both White and Black youth. They examined 77 White and 82 Black adolescents (67 male and 92 female) who had an average age of 12.2. They focused on students who were considered both successful and at risk and employed the Multigroup Ethnic Identity Measure (MEIM) to assess the strength of ethnic identity, and the Children's Depression Inventory (CDI), and the Child Behavior Checklist (CBCL) to assess adjustment. Conversely, lower levels of ethnic identity predicted lower levels of social and emotional adjustment. Employing the Adolescent Discrimination Distress Index (ADDI), Racial Socialization Scale (RSS), and the Rosenberg Self-Esteem Scale (RSES), Harris-Britt, Valrie, Kurtz-Costes, and Rowley (2007) examined how racial socialization messages can protect African American adolescents from perceived racial discrimination. In their study of 128 African American eighth-grade students (68 girls and 60 boys) from two public schools in the southeastern United States, they found that although racial pride did not predict self-esteem, other forms of racial socialization (i.e., understanding of cultural history, expectations of possible bias/discrimination) could positively predict higher levels of self-esteem. One area of recent interest is the mediating/moderating influences of racial identity on selfesteem. Eccles, Wong, and Peck (2006) examined how ethnic identity could mediate the effects of racial discrimination in African American adolescents. They drew their results from the 1990s-era Maryland Adolescent Development in Context Study (MADICS), which examined a middle-class community with 51% African American and 43% European American households that have relatively equal income/political influence in the community. They focused on 1,480 middle-school-age adolescents over a 2-year period (seventh and eighth grade). The researchers measured the self-reported experiences of discrimination at school as well as students' future expectations of discrimination and how these experiences influenced their views of and performance in school. Also measured were the students' self-reported feelings of “positive connections with their ethnic group” (their way of assessing ethnic identity). Regression analysis revealed that a positive connection to ethnic group predicted grade point average during the eighth-grade year. Indeed, the measured effect was approximately equal to the negative effect of experiencing racial discrimination, suggesting that a positive relationship to ethnic identity can counter, protect, or minimize the effects of perceived racial discrimination on academic motivation and achievement in African American adolescents. Gaylord-Harden, Ragsdale, Mandara, Richards, and Petersen (2007) explored how ethnic identity and self-esteem function as mediators between social support and internalizing symptoms such as anxiety and depression. A sample of 277 (85 males, 142 females) African American middle-school students from high-crime rate neighborhoods in Chicago was used. The researchers utilized the MEIM, Survey of Children's Social Support (SOCSS), Experience Sampling Method, and the State-Trait Anxiety Inventory for Children (STAIC) to gather results about the sample's ethnic identity, perceived social support, self-esteem, and anxiety. They found that the strongest predictor of depressive symptoms and anxiety was perceived social support. Ethnic identity and self-esteem seemed to account for males' perceived support and

females' self-esteem respectively. The gender finding is important and should suggest research questions that further differentiate the experiences of Black boys and girls. Oney, Cole, and Sellers (2011) explored how racial identity and gender moderated self-esteem and body image for African American men and women. They drew on a sample of 425 college students (109 male and 316 female) from three regions of the country (East Coast, Southeast, Midwest). Utilizing several subsets of the Multidimensional Inventory of Black Identity (MIBI), they found that for those whose race was less central to their identities, body dissatisfaction was related to lower self-esteem. For those subjects who had high private regard (of their identities) and low body dissatisfaction tended to have higher self-esteem. Those who had low public regard and high body dissatisfaction displayed lower levels of selfesteem. The researchers also looked at assimilation ideology (i.e., emphasis on similarities between African Americans and Western society) on self-esteem; however, no significant relationship was found. Not unexpectedly, the connection between body image and self-esteem was higher for females than for males. The researchers also looked at assimilation ideology (i.e., emphasis on similarities between African Americans and Western society) on selfesteem; however, no significant relationship was found. Not surprisingly, given the higher salience of body image particularly for females, the connection between body image and selfesteem was higher for females than for males. Some research has shown that the differences in self-esteem among Whites and Blacks can be attributed to the design of self-esteem reporting methods (Bachman & O'Malley, 1984). Bachman and O'Malley (1984) showed that Black adolescents were more likely to report more extreme levels of self-esteem than Whites when using incremental-based (1–5) methods. When these differences were accounted for, the gap between levels of self-esteem were not as acute. Some have explored the relationship between parental influence and social/personal selfesteem in Black adolescents (Constantine & Blackmon, 2002). The work illustrates the more complex parental tasks required of Black socializing agents and contributes an understanding of the John Henryism well-being challenge referenced previously. In fact, Constantine and Blackmon (2002) found that parental messages that reinforced pride and value in African American culture were positively associated with self-esteem among Blacks and their peers. However, those who emphasized the preference of White/majority institutions and culture were negatively associated with school self-esteem in Black adolescents. Research has shown that for U.S. high school students, self-esteem is comparatively high, with Blacks scoring the highest, followed by Whites, Hispanics, and then Asians (Bachman, O'Malley, FreedmanDoan, Trzesniewski, & Donnellan, 2011). Slightly higher levels of self-esteem were reported by males compared to females. Tynes, Umaña-Taylor, Rose, Lin, & Anderson (2012) explored the moderating influences of racial identity and self-esteem on the negative effects of online racial discrimination (i.e., depression and anxiety). They surveyed 125 African Americans ages 14 to 19 (55% female) from three urban high schools in the Midwest and compiled composite scores measuring online racial discrimination, ethnic identity, depressive symptoms, anxiety, and self-esteem. The researchers found that the negative effects were significantly minimized for those who reported higher levels of racial identification and self-esteem and observed that males experienced

significantly less discrimination than females. Researchers operating within a PYD perspective have tended to criticize the approaches of the risk and resiliency models, given their emphasis on assets mediating/protecting at-risk youth from negative outcomes rather than an emphasis on the importance of positive assets (Evans et al., 2012). They point toward work that better measures the association between racial socialization messages such as preparation for bias with positive racial identity. They believe that this approach is more conducive to PYD concepts such as confidence, competence, and connection (Murry et al., 2009). Lusk, Taylor, Nanney, and Austin (2010) explored the relationship between racial identity and self-esteem for Black/White biracial individuals. The average age of the sample in this study was higher than the age in other works cited in this review (23.62 years), but this study has been included to provide a wider perspective and because of the dearth of work on biracial identity and self-esteem in adolescents. Lusk et al. focused on 74 self-identified biracial individuals drawn from online multicultural websites as well as university campuses, which clearly limits the generalizability of their findings to more affluent/privileged populations. They utilized Rockquemore and Brunsma's racial self-identity measure (2002), MEIM and RSES and the Center for Epidemiological Studies Depression Scale (CES-D) to determine levels of emotional well-being. They found that the individuals in their study self-identified first as biracial (“border identity”) and then a singular one (i.e., White/Black). This was in keeping with Rockquemore and Brunsma's (2002) original work on biracial identity. Lusk et al. did note, however, that more participants identify differently depending on context/circumstance. The overall results of the study confirmed previous findings regarding the positive relationship between racial identity and self-esteem as well as the negative correlation between identity and depression. Additionally, they found that those who identified with border identities exhibited higher levels of self-esteem than those who identified with singular or “transcendent” options. The literature seems to point toward a generally positive association between racial identity and self-esteem, at least in regard to African American youth. Disagreements between researchers focus on conceptual differences relating to how to measure racial identity (i.e., pride of the group or affirmation of self) and whether this identity promotes/protects minority youth from stressors or whether that identity serves as a positive asset in and of itself. However, certainly the problem of colorism plays a very important role in the lives of youth of color. The research suggests that group membership is salient and, is particularly evident, when associated with contextual experiences. Given historical, underacknowledged, and generally ignored historical events and evolved social traditions, color matters.

Research Acknowledging Colorism To address the recent literature surrounding the phenomenon known as colorism, a definition of the term is critical. Burke (2008) describes it as “the allocation of privilege and disadvantage according to the lightness or darkness of one's skin” (p. 17). The literature on colorism is vast, beginning with the classic Clark Doll Studies (Clark & Clark, 1940), which showed that Black

and White children demonstrated a consistent preference for a White doll over a Black one (reviews are available that include at least 50 years; see Cross, 1991, 2003; Spencer, 1982a, 1976; Spencer & Markstrom-Adams, 1990). For a good overview of some of the principal contributions from the expansive literature on doll studies, see Fegley, Spencer, Goss, Harpalani, and Charles (2008). More recent treatments have expanded, revised, and altered aspects of Clark's original work with some exploring how such preference (i.e., given skin color) might be altered through incentives (Jordan & Hernandez-Reif, 2009). Researchers explored this bias for other racial groups with Hispanic children also demonstrating a bias toward lighter skin (Kaufman & Wiese, 2012), although some have been skeptical of the reliability of traditional doll studies (Stokes-Guinan, 2011). Studies have also shown that Black children are more likely to remember narratives that depict darker-skinned African Americans in a negative or stereotypical light (Averhart & Bigler, 1997). Overall, these studies have tended to reinforce the original results of the Clark study with incentives and technological changes having limited impact on changing child and adolescent attitudes toward skin color (Fegley et al., 2008). Despite these challenges, research into the effects of colorism on children and adolescents continues to be a dominant area of inquiry with particular emphasis on the negative impact of privilege on their development and life experiences. Scholars often depict colorism as a social ideology bound up with the institutions of public life, such as school and the workplace. The family, it would seem, serves as a place of refuge from which its members might contest dominant racial narratives. Some research has shown, however, that Black families often reinforce colorism by privileging lighter-skinned members (McDonald, 2006; Russell, Wilson, & Hall, 1993; Wilder & Cain, 2011). Russell et al. (1993) suggested that a family's preoccupation with skin color begins at the time of the child's birth. McDonald (2006) explored a case study where a woman born with darker skin than her mother and siblings was abused and neglected as a result of her mother believing that she had “lowered” the family's status. Wilder and Cain (2011) showed that among Black women, maternal figures often play a central role in shaping Black children's attitudes about skin tone. Many of these women claim that they learned at a young age to associate darker skin with ugliness and lighter skin with beauty. One area of research that remains particularly voluminous is in the impact of colorism on the social mobility of African Americans in United States. Many investigators examined the relationship between skin color and educational attainment and economic success in both contemporary and historical contexts (Bodenhorn, 2006; Bodenhorn & Ruebeck, 2005; Goldsmith, Hamilton, & Darity, 2006; Hannon, DeFina, & Bruch, 2013; M. Hughes & Hertel, 1990; Keith & Herring, 1991; Ryabov, 2013). These studies have found that even when accounting for parental variables, lighter-skinned Blacks are more likely to obtain higher levels of education and pursue more prestigious and lucrative professions than those who have darker skin. In addition to professional and education success, some researchers have focused on more subjective evaluations, such as personal attractiveness (M. Hill, 2002; Wade & Bielitz, 2005).

Drawing from the National Survey of Black Americans, Hill found that skin color had a greater impact on the perceived attractiveness of women than on men with lighter skin being seen as a more attractive quality. Wade and Bielitz (2005) examined perceptions across gender and race and found that African American women with lighter skin were rated higher than lighterskinned African American men on intelligence and parenting. Women generally rated darkskinned individuals higher than men did on intelligence. Self-esteem and self-efficacy have been important topics within the literature of colorism with some exploration of the differences affecting male and females (M. Hill, 2002; Thompson & Keith, 2001). Thompson and Keith (2001) demonstrated that skin color remains an important indicator for self-esteem in women but not in men and for self-efficacy in men but not in women. They also maintained that, for women, the power of colorism was strongly associated with the individual's social class with working-class darker-skinned women having lower self-esteem than dark-skinned women in higher social classes. Elmore (2009) explored the relationship between self-perceptions of skin tone and the processes of identity formation in adolescents. Other scholars have shown that young women who embrace an African American standard of beauty (reflecting a higher level of self-esteem) exhibit lower levels of substance use than those who embrace White standards of beauty (Wallace, Townsend, Glasgow, & Ojie, 2011). As suggested previously, in light of Barack Obama's election as the first African American President of the United States, some scholars have examined the impact of colorism on political behavior (Weaver, 2012). Weaver (2012) demonstrated the limitations of traditional discussions of race and politics and argued that skin tone and hue play a far more pivotal, however subtle, role in shaping electoral outcomes in the United States. The implications of colorism for immigrants and workplace discrimination are examined in Hersch (2011). In accounting for length of residency, English-language competency, and education, Hersch found that lighter-skinned immigrants earn 16% to 23% more on average than darker-skinned persons. The findings are important, given the role of socioeconomic status as a source of support and protection for families and children. This phenomena is not unique to the United States. There have been a number of efforts to explore the implications of colorism around the world. (For Japan, see Wagatsuma, 1968; for India, see Beteille, 1968; for dark-skinned Puerto Ricans, see Hall & Julian Samora Research Institute, 1987; for Latin America, see Russell, Wilson, & Hall, 1993.) Ferguson and Cramer (2007) showed that rural Jamaican children who self-identify and idealize skin type as White exhibited higher levels of self-esteem. Some scholars have taken up the idea of “economies of color” in examining the pigmentocracy underlying racial liberalism globally (Glenn, 2009). Charles (2011) studied how Jamaicans have historically practiced skin bleaching in order to enhance sexual attractiveness. In a survey of self-reported health and racial categories in selected Canadian cities, Veenstra (2011) found that darker-skinned Blacks were more likely than lighter-skinned ones to report poorer health outcomes. The findings of this research point toward a tendency to demonstrate preference for lighterskinned and White phenotypes. However, some research has shown a more complex dynamic where Whiteness alone is not treated as a privileged category (Quiros & Dawson, 2013).

Quiros and Dawson (2013) demonstrated how among Latino populations, although lighter skin is seen as being a more positive feature, being perceived as White does not carry the same level of prestige. The complexities of colorism therefore do not appear to be as uniformly internalized as previously believed. Other examples of research within this “revisionist” camp have shown that skin color does not necessarily exacerbate racial discrimination outside of the family but that there is a correlation between skin tone and family dynamics (Landor et al., 2013). This research maintains that parents will demonstrate preferential treatment to their children based on skin tone. Specific gender variations were noted, with dark-skinned males and light-skinned females receiving “higher-quality parenting” and “racial socialization” within the family. Colorism and Achievement Dupree, Spencer, and Fegley (2007) tackled stereotypical assumptions about (relatively) low intellectual ability among African Americans and Hispanics. They reported research findings which suggest that many high-achieving male and female adolescents of color have self-schema that include positive racial identities. At the same time there are also high-achieving adolescents of color who do not. Their study investigated skin color preference of highachieving Asian, Black, and Hispanic adolescents and was framed to indicate cognitive coping outcomes in response to skin color stereotypes and Eurocentric standards of beauty. In addition, relationships among different dimensions of youth self-schema (i.e., skin color preference, academic and social self-esteem, and body esteem) were explored. Place of origin and race-ethnicity were used as proxies for the different culturally based skin color connotations to which youth are exposed. Results suggest that, in interaction with race-ethnicity or place of origin, preference for lighter skin predicts academic and social self-esteem for high-achieving adolescent females; however, this was not the finding for high-achieving adolescent males. In interaction with place of origin, preference for lighter skin was a predictor of body esteem for both high-achieving males and females. These gender differences suggest gender-specific self-schema for high-achieving youth of color. Skin Color and Well-Being High academically performing adolescents of color negotiate multiple identities (e.g., gender, race, ethnicity, academic, etc.) (Bonner, Lewis, Bowman-Perrott, Hill-Jackson, & James, 2009; A. Graham & Anderson, 2008). The very idea that youth “negotiate” multiple identities suggests that there can be a variety of combinations of identity outcomes. These processes can differ in how central gender, race, ethnicity, and academics are to their identities. Research targeting high-achieving African American students provides many examples of such youth who are both high achieving and maintain positive racial identities (Bonner et al., 2009; D. J. Carter, 2007; Carter Andrews, 2009; A. Graham & Anderson, 2008; Harper, 2006; Spencer, Noll, Stoltzfus, & Harpalani, 2001). Some of these same studies demonstrate that these highachieving youth not only hold race central to their identities but (a) they also have a critical understanding of the structural or systemic forces that work against the positive development and achievement of racial and ethnic minorities, and (b) they seek to actively resist these

forces (Bonner et al., 2009; D. J. Carter, 2007; Carter Andrews, 2009; A. Graham & Anderson, 2008). Salient about the findings is that youth are not resisting or opposing a culture of achievement. They are resisting stereotypical conceptions of their racial or ethnic group. Specifically, they resist assumptions that they are underachievers and assumptions about the superiority of members of the majority culture. Yet, also critical to acknowledge is that these data do not preclude the possibility that there are high-achieving youth who do have negative racial identities. It is more likely that there are many different identity profiles that reflect individual responses to proximal experiences of race and ethnicity. In the United States, a critical aspect of our individual experiences of race is visibility (i.e., what we look like, who we appear to be). Skin color influences visibility as it is one critical characteristic used to determine race, along with facial features, hair texture, and body type (e.g., Stepanova & Strube, 2009). Carter Andrews (2009) focuses on the skin color dissonance of high-achieving youth of color as one indication of how these youth are negotiating their racial identities. It represents another step forward by operationalizing racial identity in terms of actual physical status—the consonancedissonance between one's actual skin color and one's preferred skin color. The study also explores how the skin color dissonance of high-achieving youth of color—as representative of cognitive coping outcomes in response to skin color stereotypes and Eurocentric standards of beauty—relate to other dimensions of their self-schema (i.e., academic and social self-esteem and body esteem). Culturally Based Color Connotations and Skin Color Consonance-Dissonance In many places, there exist skin color connotations that are qualitatively different from those in the mainland United States (Benson, 2006). Group differences may exist but not necessarily along the lines of White versus non-White. For instance, in the commonwealth of Puerto Rico, ethnic identity takes precedence over racial identity, and the caste system is a function of the culturally determined meanings given to SES, facial features, and hair texture in addition to skin color (Alarcon, Szalacha, Erkut, Fields, & García Coll, 2000; Benson, 2006; López, 2008; Montalvo, 2004). Under such conditions, culture may have qualitatively different influences on skin color preference. Yet cross-cultural experiences, such as immigration from one country to another or regular travel between one country and another, can make the symbolic, culturally determined role of skin color much more explicit. That is, if one recognizes that the same skin color is perceived and responded to differently depending on the cultural context in which one resides, it becomes much more obvious that skin color connotations are culturally determined and not necessarily the norm for all. However, as it relates to skin color, many places share culturally based color connotations that are similar to those of the mainland United States. There are places where—as in the United States—lighter or White skin is more valued (R. Frank, Akresh, & Lu, 2010; Uhlmann, Dasgupta, Elgueta, Greenwald, & Swanson, 2002; Ullrich, 2010). Accordingly, there would not necessarily be dissonance-creating immigration experiences that would raise awareness of the arbitrary and symbolic character of skin color connotations.

Individual experiences of skin color can raise awareness of the complex relationships among race, culture, and class for immigrants to the United States when, based on skin color, immigrants can be mistaken for nonimmigrants. For instance, in the United States, darkerskinned Hispanics with relatively coarse hair or Caribbeans of African descent may be treated like African Americans based solely on their physical appearance. Although they may originate from different cultural backgrounds, their mutual experiences of prejudice and discrimination in the United States work to acculturate these immigrants to the disfavored or stigmatized minority status of U.S.-born African Americans (Benson, 2006; R. Frank et al., 2010). Immigrant status can complicate self-perceptions based on skin color. Telzer and VazquezGarcia (2009) found that, in their study, U.S.-born Latinas with darker skin did not want to change their skin color, did not feel less attractive, and did not have low self-esteem. In contrast, the immigrant Latinas in their study wanted their skin color to be lighter, felt less attractive, and had lower self-esteem. These findings suggest that place of origin must be taken into account when conducting research examining skin color. As it relates to European standards of beauty (e.g., a preference for lighter or White skin), research by Sahay and Piran (1997) suggested that the farther away from the European ideal individuals are (e.g., darker-complexioned South Asian Canadian females), the less likely they are to endorse a European ideal (i.e., less likely to prefer lighter skin). Likewise, the closer individuals are to the European ideal (e.g., lighter- or medium-complexioned South Asian Canadian females), the more likely they are to be influenced by this ideal (i.e., more likely to prefer lighter skin). In fact, Fegley, Spencer, Goss, Harpalani, and Charles (2008) reported, in their study of adolescents of color in the northeastern United States, that light-skinned adolescents were just as likely to prefer their own skin color as they were to prefer another skin color. In contrast, medium-complexioned adolescents were more likely than their lightand dark-skinned peers to prefer a skin color other than their own. Likewise, dark-skinned adolescents were more likely than their light- and medium-complexioned peers to prefer their own skin color (Fegley et al., 2008). Skin Color as a Signifier of Attractiveness: The Complex Relationships between Skin Color and Body Esteem Greater satisfaction with one's skin color in later stages of development (i.e., adolescence and young adulthood) is associated with higher self-esteem and body esteem (Falconer & Neville, 2000; Telzer & Vazquez-Garcia, 2009). In her study of ethnicity and body image, Altabe (1998) found that, in general, Asian Americans and African Americans held similar perceptions of the cultural ideal for skin color (i.e., lighter skin color is believed to be desired by other men and women). Relatively speaking, Asian Americans had the highest desire for lighter skin followed by African Americans. Hispanic Americans had the least dissatisfaction with skin color and valued darker skin or wanted darker skin color. However, Hispanic Americans showed more body dissatisfaction than Asians or African Americans. Interestingly, African Americans had higher self-ratings of attractiveness than Hispanic Americans. Likewise, Hispanics had higher self-ratings of attractiveness than both Caucasian Americans and Asian Americans. But, when asked to rate the importance of appearance to their self-

esteem, all groups scored higher than the Asian Americans. In a study based in Canada, Sahay and Piran (1997) found that South Asian Canadians showed the least skin color satisfaction compared to European Canadian females and that light- and medium-complexioned South Asian Canadian females reported the lowest body satisfaction. Thus, a desire for lighter skin color likely reflects a devaluing of one's own skin color and a desire to identify with individuals who have lighter skin color. These findings reflect the complexity of identity negotiations even as they relate to physical appearance and perceived attractiveness. Adolescents' Negotiation of Skin Color Consonance-Dissonance as a Coping Response Given the topic's long-term silencing in the social science literature except for inferred assumptions about its psychopathological character for Blacks (for a review, see Guthrie, 1976), attentiveness to the proactive coping of youth with dissonance-producing contexts has been ignored. It might be that in the United States, wanting their skin to be lighter or darker is psychologically important for adolescents of color. Specifically, in race-based societies like the United States, such cognitions can represent attempts to adapt to the perceived cultural standards. The cognitions represent coping responses. As suggested by the theoretical grounding provided by the phenomenological variant of ecological systems theory (PVEST), identity processes are coping processes (Spencer, 2006; Spencer, Fegley, & Harpalani, 2003). More specifically, identity outcomes emerge from consistent coping behaviors (i.e., particularly patterned responses). Consequently, resulting identities serve an organizing function as they influence one's thoughts and behaviors (i.e., determining which thoughts and behaviors that do and do not fit with my identity). Spencer (2006) made a distinction between reactive coping and stable coping. Whereas reactive coping represents responses to specific and immediate challenges, stable coping represents general responsive strategies that are engaged consistently without significant regard for the specific character or nature of challenges encountered (e.g., having health consciousness as a significant aspect of one's identity, the automatic response to “pass” on the offer of a cigarette does not require conscious consideration). Particular challenges may be encountered quite regularly; accordingly, as a response, specific coping strategies are responsively and consistently employed (i.e., thus representing stable coping strategies); the patterned response contributes to and aids in maintaining an emergent identity status. The resulting identity serves as a filter for day-to-day (potentially and ideally healthy) functioning. There will be certain thoughts and behaviors that are in keeping with one's identity and, at the same time, other thoughts and behaviors that are not. A PVEST interpretation would suggest that skin color preference represents (a) a coping response to pervasive and consistent exposure to culture-specific color connotations and racial or ethnic stereotypes and (b) an identity outcome to the extent that it suggests the degree of acceptance of one's own skin color. For instance, when an individual does not fit the skin color ideal within a particular cultural context, skin color preference signifies his or her racial ideology (i.e., one's acceptance or rejection of a culture's overarching color connotations). A preference for lighter skin reflects a desire to be perceived as more similar to the other individuals who have light or White skin and to benefit from all the advantages and privileges afforded to those with light or White skin (i.e., perceived and treated as more attractive,

smarter, morally upright, etc). For instance, research supports the notion that people who are perceived as more attractive (based on skin color or otherwise) are also perceived as smarter and more likely to be successful. In some instances, lighter skin is associated with more successful life outcomes, while darker skin is associated with less successful life outcomes (i.e., personal earnings) (R. Frank et al., 2010; Harrison & Thomas, 2009; Hunter, 2002; Kiang & Takeuchi, 2009; Townsend, Thomas, Neilands, & Jackson, 2010). Gendered Differences in How High-Achieving Adolescents Negotiate Multiple Identities As suggested by data, there are gender differences in the relationships between skin color and other dimensions of self-schema. For instance, as skin tone relates to self-perceptions, Thompson and Keith (2001) found that, for both men and women, the lighter the skin tone, the higher the sense of self-efficacy. However, the effect was more pronounced for men in the study. (Of course, these findings may be somewhat different for youth, given the study's focus on adult participants. Other research suggests that females are more likely to desire lighter skin color than males (e.g., Fegley et al., 2008; M. Hill, 2002; Wade & Bielitz, 2005). In relation to perceptions by others, Wade and Bielitz (2005) found that light-skinned African American females are perceived as higher in intelligence, a relationship that does not necessarily hold for light-skinned African American men. Together, these findings suggest that gendered experiences may influence the role of skin color for the identities of adolescents of color. Specifically, gender differences in their experiences may lead male and female African American adolescents to think about skin color—as a signifier of race—very differently. Overall, findings from African American youth suggest complex relationship and individual differences—including gender patterns—that index complex relationships among self, context, and identity as coping processes. In particular, the skin color findings suggest very interesting, nuanced, and overall complex color-associated patterns. More informed, theory-based, and rigorous research is needed for testing and integrating (1) the historical analyses by D. B. Davis (1966); (2) the critical perspective by Harris (1995), which hypothesizes the representation of “Whiteness as property”; (3) critical race viewpoints of Roediger (1991) alluding to the “wages of Whiteness” (i.e., as affirming and supporting myriad benefits); and (4) the knapsack of supports and privileges suggested by McIntosh (1989). Along with others viewpoints, these are not typically woven into the constructs considered in social science research or reflected in the research methodologies and designs embraced. Most important, for African American youth, their self-identity formation generally does not represent developmental processes linked with their identifiability—skin color—in light of their progressively complex meaning-making capacity (i.e., given cognitive maturation irrespective of its expression in achievement-assessing contexts). The final sections of this chapter describe in more detail both the methodological challenges and the opportunities possible. They also consider the conclusions and inferences possible from a social science perspective that acknowledges the impact of a 150-year historical backdrop in which color continues to matter.

Methodological Considerations In theorizing about individuals' ability to resist and overcome risks, Prilleltensky (2008) asserted that power is always reflected in political processes which create conditions impacting psychological processes. We have framed the historical influences and research on vulnerability and resilience within this orientation. These factors provide not only the context in which development occurs but also the one in which research is conducted. The sociopolitical nature of most scholarship on African Americans requires significant consideration of issues involved in conducting research. “Central to a discussion of procedural issues in minority research is the awareness that research on minority issues often has perpetuated stereotypes, reflected prevailing biases, and assumed that personal characteristics of research participants reflect causal connections to specific outcomes” (Schwartz et al., 2014, p. 62). With that assertion, the next review presents conceptual and procedural issues related to research on resilience among African American youth.

Conceptual Issues The study of resilience, with a focus on processes through which it develops, requires conceptual framing that offers criteria for defining and examining the processes. The framing defines the scope of risk in understanding intrapersonal, interpersonal, and contextual factors contributing to unexpected trajectories, given the magnitude of vulnerability. Risks are generally well defined, given their explicit associations with adverse outcomes, which include issues that are school related (i.e., academic failure and school dropout), health related (i.e., obesity and sexual behaviors), justice system related (i.e., various forms of delinquency), and mental health focused (i.e., anxiety, depression). Understanding mechanisms by which protective factors mitigate these risks has been a primary focus of resilience research. The Scope of Risk Resilience research evolved from examining trajectories associated with youth experiencing positive outcomes despite numerous compromising risks. This shift expanded historical research from a pathologizing orientation to including considerations for strength-based studies (Zimmerman et al., 2013). Over the past two decades, complex research designs have examined varying forms of vulnerability and risk in exploring differences in developmental trajectories and levels of need to be mitigated (Jung & Wickrama, 2008; Sohoni & Soporito, 2009). Consistent with the methodological and conceptual challenges associated with late twentieth-century and early twenty-first-century social science research traditions alluded to, importantly and unfortunately, few focus on African American youth. Arrington and Wilson (2000) have suggested differentiating “risk” from “vulnerability.” They asserted that the latter represents individual factors, in contrast to the former reflecting group experiences. Resilience theorizing, however, continues to broadly conceptualize risks as contributing to significant vulnerability for youth individually and collectively. Resilience, in essence, represents the individual experience of risk and the individual impact of vulnerability. Arrington and Wilson, however, introduced a consideration for explicitly acknowledging group

vulnerabilities as a condition of individual risks. Based on the dual-axis model of vulnerability (Spencer, 2006; Spencer, Harpalani, Cassidy, et al., 2006), all youth experience varying risks and levels of support. Many studies explore factors associated with the most vulnerable youth: high risk, low supports. The scope of risk frequently considers poverty as a contextual condition indicative of restricted resource accessibility, but poverty is also recognized as an inaccurate assessment for determining youth success. As a protective or support mechanism in mitigating risks associated with poverty, numerous studies confirm the impact of a trusting and caring adult as a significant, contributing factor toward youth resilience (see Zimmerman et al., 2013). The consistency of this finding became the basis for mentoring programs and a central component of youth-focused intervention projects more generally (DuBois & Karcher, 2005; Herrera, Grossman, Kauh, Feldman, & McMaken, 2007). This body of research highlights the effects of caring adults throughout adolescence and into emerging adulthood, where many colleges and universities have incorporated mentoring urban youth into components of their service learning offerings. To understand attributes of particularly effective relationships, some studies have examined the individual histories and experiences of significant adults to capture their moderating impact on youth outcomes. These include discrimination experiences, racial socialization, earlier school experiences, substance abuse, and personal traumatic experiences. Acknowledging the direct and indirect impact of significant adults' histories is comparable to accounting for the impact of cultural trauma or sociohistorical influences (discussed later) in creating conditions of risk for contemporary youth. Collaboration among families, schools, and community-based organizations are critical to building supports for the future trajectory of youth (Broussard, Mosely-Howard, & Roychoudhury, 2006). The greater the social support, the more competent and motivated students are in their educational abilities. As such, the guidance and care demonstrated through mentorship heightens willingness for school engagement as a basis for academic achievement and in deterring decisions to leave school (Darensbourg & Blake, 2013). Although research documents the positive impact of mentoring relationships, little research has been done on the nature of these relationships and the youth targeted to receive mentorship. For example, Monroe (2013) found that colorism impacts both youth opportunities for educational opportunities and also selection as a mentee in nonschool programs. Although there is a clear trend favoring lighter youth (and females over males), research regarding criteria for selecting youth participants remain unclear. These approaches on interpersonal dynamics also draw from research on the impact of close peer relationships associated with resilience during adolescence (Masten et al., 1999; Morrissey & Werner-Wilson, 2005). Youth without trusting and stable relationships can experience difficulties navigating contextual risks that strain their limited coping abilities. There may, however, be critical or sensitive periods during which experiences influence the subsequent interpretation and response to future events of youth (see Swanson, 2010). A youth's history of unstable relationships, for example, has implications for social and emotional development and identity formation throughout the life course (e.g., ability to get along and work well with peers in school).

Phenomenological research has shown the value of accounting for the youth perspective in designing empirical studies but also for interpreting findings within a developmental orientation. This is particularly relevant given the shift in cognitive abilities experienced during adolescence related to self-appraisals and the abstract processing related to the assessment of risks and supports perceived as available. Work acknowledging the impact of poverty, for example, consistently indicates the limited opportunities available for exploration that support competence formation, self-appraisals, and identity development. It is misleading, however, to suggest that youth interpret their impoverished conditions as a risk or limitation (i.e., reports of low aspirations, or a narrow perspective of the future, by youth living in poverty is not indicative of them interpreting their future based on their family's income status). Similarly, it is misleading to suggest that mentoring relationships are supportive without also acknowledging the relevance of the relationship to the youth. The scope of risk is subsequently related to a clear conceptualization regarding the form of risk and also the historical nature of supports (i.e., stability) relevant to interpretation of experiences and supports by youth. Attributes of Resilience Other factors associated with resilience include individual attributes: an area of emerging or demonstrated competence, a positive reference group–oriented identity, and a future orientation (Masten et al., 1999; Mello & Swanson, 2007). As noted in the introduction, normalization of Whiteness remains a major problem in research agendas targeting or including African American youth and may be one of the major impediments to gaining a full understanding of development among diverse youth. White middle-class youth continue to represent the standard for normative outcomes (i.e., competence) with minimal reference to their sociohistorical status or their status as a racialized group (e.g., enjoying the “wages of Whiteness” (see Roediger, 1991). Helms' (1990) and McIntoch's (1989) work prompted examination of White racial identity and White privilege, respectively. The expansion of Whiteness and White identity studies was a major theme in research for a few years preceding conceptual considerations regarding White privileging status. Although studies have examined the impact of White identity on stereotype beliefs (R. T. Carter, Helms, & Juby, 2004), very few studies have examined White identity or privilege as a source of vulnerability (i.e., the intersection of available supports and risks) on the mental health of White youth when they are challenged and on their developmental trajectories. Among the few that approximate these issues is a study that examines the “psychosocial costs” of racial attitudes for Whites. Although it examines the within-group impact of attitudes, it does not assess the influence of identity or unexamined privilege (Spanierman, Poteat, Beer, & Armstrong, 2006). Although years have passed since frameworks exploring White privilege were presented, little empirical work has examined Whites as one of many diverse groups; rather Whites continue to represent a standard by which others are judged. Sociohistorical analyses regarding these issues remain empirically unaddressed. Much research documents the impact of experiencing discrimination, but significantly less research focuses on being a person of privilege. As an attribute of resilience, significant effort has been expended raising and integrating cultural competence into activities where adults have direct interaction with youth and impact their

experiences. Assessments of cultural awareness and cultural knowledge suggest the significance of others' understanding (and subsequent acceptance) of cultural differences as normative rather than pathological. This orientation, while relevant in attempting to broaden perspectives on differences and mitigate stereotypes, has not resulted in improved outcomes associated with resilience; in fact, it reinforced an “othering” perspective (see Harris, 1995). Training on cultural competence highlights the values different groups possess and generally introduces the implications of group history on the cultural strengths and sources of collective coping that influences current experiences. Professional training, however, seldom addresses the inherent biases of those being trained; nor does it provide sufficient time for the long-term engagement necessary to examine, challenge, and create new perspectives (Fouad et al., 2009; Hatcher et al., 2013; Rogers, 2006). For example, on average, a degree program requires one semester focusing on cultural influences relevant to individuals' development and lived experiences; only 1 to 4 hours per year are required through continuing education postgraduate training and licensing in select disciplines. Although culturally based training is seldom an explicit requirement in research-focused graduate programs, it is relevant in conceptualizing and assessing how empirically examined proximal relationships are studied. Even less frequently considered is the limited cultural perspective (i.e., ethnocentric) behind the development of youth-based policies intended to address issues related to African American youth and their families. Although there remains substantial work to understand the influence of privilege as a source of significant support for youth with few known risks, there has been tremendous effort to understand how various groups live culturally. García Coll and Magnuson (1999) suggested the need to acknowledge cultural factors as an active part of the changes and influences on developmental processes… This new treatment of culture reflects the need to “unpackage” culture, that is, to consider culture as a multidimensional evolving source of influence on developmental processes within any culture rather than as a monolithic nominal variable. The unpacking of culture also requires researchers to operationalize culture in concrete variables. (p. 3) This focus in research on exploring and understanding cultural factors impacting development has been particularly true for many immigrant families and youth, although the degree of within-group diversity continues to be minimally acknowledged even for them (see Nakamura, Tummala-Narra, & Zárate, 2013; special issue on immigration status and within-group assessment considerations). Even so, less attention has been paid to cultural and historical factors shaping experiences and outcomes for African Americans that also considers their developmental needs. Universal versus Group Dynamics Tensions exist when conceptualizing intervention programs regarding the extent to which research studies exploring resilience from a developmental orientation are focused on universal versus group-specific factors. The perception that factors contributing to resilience are universally constant ignores the sociohistorical conditions that influenced the presence and

type of vulnerabilities youth must navigate. As resilience research expands to better understand the broad factors related to positive outcomes for youth experiencing multiple risks, it is important to continue examining unique group differences as well. It is interesting to note that significant research has been conducted on within-group processes, outcomes, and trajectories, especially for African American and Latino youth (Cedeno, Elias, Kelly, & Chu, 2010; Edwards, & Romero, 2008; Gonzalez, Stein, Kiang, & Cupito, 2014; Pina-Watson, Ojeda, Castellon, & Dornhecker, 2013; Stevenson & Arrington, 2009; Travis & Leech, 2014). Findings from such studies, however, are utilized infrequently in studies involving diverse racial and ethnic groups. In searching empirical studies on research databases (i.e., PsycInfo) from 2000 to 2010, frequently cited literature reviews do not often reference group-specific studies. Although this is understandable, given a focus on universal attributes, a general orientation toward excluding group-relevant research contributes to designs and interpretations of results reflecting a status quo orientation. Much of the multigroup research on resilience acknowledges group variability in reviewing the implications and generalizability of findings, but often it is limited to interpretations of findings within a universal orientation. Exploring similarities in parallel processes offers explanations and possible confirmation of universal trajectories potentially associated with developmental trajectories; considering implications associated with different processes and outcomes, however—both individual cultural factors and positionality of the group within broader societal experiences—contributes to greater interpretive understanding of group-differentiated experiences. Examining within-group variability also receives attention as investigators increasingly attempt to differentiate experiences of native and immigrant African Americans. Due to different sociohistorical conditions, immigrant youth have a cultural reference that extends beyond American borders. Ogbu (1978; Ogbu & Simons, 1998) identified these groups as voluntary (immigrant) and involuntary (nonimmigrant) minorities to acknowledge the influence of their cultural histories within the United States. This classification was based on reasons for the different groups of African descent being in the United States and the involvement of White Americans in defining their status. Although identifying with an immigrant history can provide a specific source of support (i.e., cultural identity associated with country of origin), immigrants share the stigmatizing experiences associated with being identified phenotypically with native-born African Americans (Caldwell, Guthrie, & Jackson, 2006; Seaton, Caldwell, Sellers, & Jackson, 2008). Although not all studies will have sample sizes sufficient for within-group analyses, it is relevant to consider the implication of the diverse experiences within groups that could result in different findings or trajectories. The situation is relevant even in cases of PYD research that aims to be inclusive (see Spencer & Spencer, 2015). SES is another area acknowledged in research relevant to within-group variability. African American youth in low-SES families experience risks directly associated with limited resource opportunities that continue to be the focus of much resilience research. Youth in middle- to high-income-status families experience risks associated with racial discrimination, for example, but the scope and impact of their vulnerability is less well understood (McNeil, Harris-McKoy, Brantley, Fincham, & Beach, 2014; Wood, Kurtz-Costes, & Copping, 2011). Lindsay (2011) confirmed persistent racial differences in academic performance when

examining White and African American youth from middle-class families with similar resource capital. Similarly, McNeil et al. (2014) asserted the need for further examination of contextual risks and related demographic factors influencing of the nature and impact of discrimination among middle-class youth. Developmental Trajectories and Transitions Although many resilience models are developmental in that they focus on adolescence as a period in the life course, they vary in terms of the periods studied (e.g., early, middle, or late) and the age ranges in longitudinal designs. Research has focused on specific areas of impact influenced by vulnerability during adolescence, however, with far less attention to late childhood, the transition into adolescence, and late adolescence as an emerging shift into adulthood. Interpretational information is compromised when accounting for ages as a study condition without a concomitant acknowledgment of developmental factors associated with the ages as well. As suggested by the historical framing provided, a history of vulnerabilities and risks that is both cumulative and multiplicative suggests the presence of developmentally comprising issues that often exist long before pubertal development begins. Most research, therefore, does not focus on late adolescence within a resilience framing as the needs associated with a history of academic and psychosocial risks, and therefore sustainable intervention factors that influence resilience, are numerous and complex. Exceptions to this are studies examining the success of African American youth who attend college with attention to their mental health and college retention (see Hunn, 2014; McGee & Spencer, 2013; McGee, Hall, & Spencer, 2015; Solorzano, Ceja, & Yosso, 2000). More work is emerging around middle to late childhood as precursors to adolescent experiences (see Swanson, 2010). Early childhood research provides a basis for understanding the impact of developmental processes in cultural awareness and self-processes as indicators of resilience. The relevance in particular of meaning making and the implications for children understanding and internalizing contextual supports was linked to their developing social cognition. A developmental pattern in the understanding of color connotations, for example, is a consequence of increases in social experiences and social cognition (Spencer, 1982b, 1985; Swanson, Cunningham, Youngblood, & Spencer, 2009). Three-year-old children of all race/ethnic backgrounds are generally Eurocentric in terms of racial attitudes, racial preferences, color concepts, and connotations. However, for Black children, an Afrocentric pattern of attitudes and preferences begins to appear at age 5 and continues through age 9. Different developmental patterns emerge when SES is taken into account. Middle-income Black children are generally Eurocentric in terms of racial attitudes, racial preferences, color concepts, and connotations at age 3, while low-income children were neutral on the same dimensions at the same age. After age 3, both low- and middle-income children showed a clear Eurocentric orientation for color concepts. However, middle-income children became progressively more Afrocentric through age 9, while low-income subjects showed a pattern of neutrality. This work by Spencer and colleagues, among others, became the basis of exploring the role of racial identity processes and the impact of discrimination experiences through adolescence as

predictive of resilience. A broader, developmental perspective of youth experiences and outcomes necessitates the assessment of processes that are precursors to identity formation: cognitive development and opportunities for “industry.” Findings on racial identity and discrimination experiences, for example, are broad in their exploration of mediating and moderating variables and assessments of discrimination (Eccles et al., 2006; Harris-Britt et al., 2007). Empirical reports should include information on the population criteria (i.e., middle school age versus high school age), measurement validity (i.e., assessing types of discrimination), and context of experience (i.e., racial demographics). Given differences in early opportunities received, cognitive shifts that occur during adolescence, and the general developmental needs and areas of vulnerability between early and late adolescence, aspects of resilience are unique at different periods. Research examining the middle school and high school periods often focus on academic outcomes, identity development, and relationships with adults and peers. Post–high school research (i.e., emerging adult years) explores issues largely related to academic experiences in college and university settings. Approaches that focus on the individual experiences and issues contributing to resilience have been necessary in understanding the impact of contextual factors on youth; they have also provided sufficient data demonstrating the extent to which unaddressed contextual factors can undermine long-term outcomes. Examinations of resilience therefore extend from the conceptual to the procedural.

Procedural Issues As others have noted, the scope and quality of methodological procedures have implications for the value of research findings (see Schwartz et al., 2014; Veale, 2006). Two major areas of consideration are briefly highlighted here: (a) recruitment and sample characteristics influencing generalizability of findings and (b) psychometrics (e.g., assessment reliabilities and normed criteria). Recruitment Considerations Research participants provide data that inform the interpretation of findings relative to the questions being examined. Characteristics of the sample also affect attrition and generalizability of the findings. Inclusion and exclusion criteria are important to note in reporting studies. Recruiting in urban schools with high free-lunch participation or community mental health facilities to address issues relevant for a broader group is exclusionary by default; students in schools with diverse SES populations are excluded, as are youth receiving mental health services paid or supplemented through private insurance. Such inclusion and exclusion approaches are based on a priori decisions but can also reflect limited time or effort to establish relationships with individuals in locations that facilitate accessibility to diverse populations. The impact, regardless of the reason, is suggestive of skewed results with limited generalizability. This issue is often addressed in discussing the implication of findings, but it is not frequently raised in considering future research strategies. It can be difficult to gain access to youth identified as low risk because they are more likely to be protected, whether having high or low supports, from research efforts perceived as exploitive (Hatchett, Holmes, Duran, & Davis, 2000). Research briefs on effective recruitment

approaches in conducting developmental research with African American youth or their families are generally not available but could broaden existing approaches used by investigators. Psychometric Considerations It is common, and often expected, for psychometric properties of new instruments to be investigated and documented. It is particularly critical to do so with instruments adapted for a developmental period or used with a population not previously tested. It is useful, although uncommon, for investigators to report the reliability of measures for each group being studied if there is limited documentation of the measure's prior use with each group. The assumption, and therefore the interpretation of results, is that all groups will interpret items (i.e., on selfreported measures) representing a construct similarly and consistently. This level of planning and reporting will enhance the likelihood of measurement strategies representing the constructs investigated. Research with diverse groups may occasionally require alternative interpretations for established constructs. The constructs of racial awareness and identity, for example, may be conceptualized as risk variables in exploring the impact of discrimination. However, youth have varying experiences of race as a risk factor related to their skin tone. African American toddlers and young children, for example, have not developed a racial identity but demonstrate awareness of color connotations, or concepts, as early as age 3. This early awareness of race and color connotation is related to the child's social cognitive ability to take the perspective or point of view of another, which becomes more differentiated in self-appraisal processes with maturation (Spencer, 1982b; Swanson et al., 2009). In essence, perceptions and stereotypes related to skin color can influence self-appraisal in conjunction with racial identity. Skin color biases pose a risk for youth of color, particularly African American youth, as skin color is the most visible and salient phenotypic feature associated with racial categorization and related biases. There is a need to examine “norms,” both in terms of attributes associated with typical development—emotional, cognitive, social—and in terms of data typically collected to assess specific constructs (e.g., connectedness, individuation, etc.). For example, in the absence of examined criteria that defines norms, there is potential for misinterpretation of developmental processes that underlie constructs in diverse groups. In using family structure as a construct of normed family patterns, Murry, Smith, and Hill (2001) reiterated the need for research to redefine family composition and family structure to more accurately capture the family supports available to African American families not represented in traditional assessments using two-parent and single-parent criteria. They noted that this construct maintains a consistent definition of single parents as representing both composition (i.e., the individuals in families) and structure (i.e., single, two-parent, intergenerational). Given the significant reporting and inclusion of family structure as a factor associated with vulnerability, greater differentiation can enhance empirical reports and account for the impact of extended family on the development of African American youth.

An understanding of racial attitudes in the context of social cognitive development suggests that some explanatory power can be lost when race (used as a categorical variable) is considered as a risk factor without additional variables that captures the experience of race (i.e., the youth's understanding of color connotations, racial attitudes, and preferences). In essence, identity development may also reflect processes by which a youth copes with her or his experiences of race (Swanson, Cunningham, & Spencer, 2003). For instance, research suggests that there is a difference between children's ability to perceive color connotations and their acceptance or internalization of those color connotations (Spencer, Harpalani, Cassidy, et al., 2006). A similar pattern is found among youth when exploring the impact of their experiences with discrimination and factors contributing to depression (Caldwell et al., 2006; Lambert, Herman, Bynum, and Ialongo, 2009; Seaton et al., 2008). Depending on the phenomena being studied, the internalization of negative stereotypes associated with being a member of a particular racial group may be a more important risk factor than belonging to the specific racial group. Social cognition among adolescents is seldom explored, yet it is a critical construct for exploring the meaning-making processes associated with self-appraisal processes and subsequent behavioral coping patterns. Contextualizing Norms Research with diverse groups also necessitates establishing new norms or reassessing the applicability of current conceptualizations. Often norms serve as benchmarks for determining whether behaviors and abilities are within the expected range versus below or above it. Without an understanding of norms for diverse groups, children from different groups may be mistakenly identified or overlooked for intervention when their behaviors or abilities are based on the degree to which expected norms are expressed. For example, there are standard criteria for classifying a child as anxious, yet there are group differences in the expression of anxiety (Morrissey & Werner-Wilson, 2005; Wiese, 2010). The limited degree to which cultural differences exist in coping expressions is a challenge. This is particularly relevant when examining a developmental period noted for heightened anxiety in addition to the criteria used for observer-rated assessment procedures to evaluate behavior problems. Clinical standards for assessing behavior problems would appear sufficient for identifying needs and subsequent treatment plans if there were not continuing concerns regarding overdiagnosing African American males for behavioral plans within schools (see Noguera, 2008). As DiBartolo and Rendon (2012) found in their review of literature examining cultural factors influencing mental health outcomes, sociocultural factors significantly affected the expression of symptoms with implications for treatment. A child exposed to neighborhood violence may, over the course of time, exhibit behavioral impulsivity more consistent with hyperactivity or conduct disorder than with anxiety. National organizations approach violence as a health risk (i.e., Centers for Disease Control) and a psychological risk (i.e., National Institute of Mental Health). The implications, however, of exposure to violence on youth resiliency has not been systematic or conclusive; the association between violence and trauma is defined quite broadly across various studies. Nevertheless, as noted, there is evidence of somatic complaints among youth who exhibit behavioral problems who have also, for example, witnessed violence. For

African American youth, little research explores behaviors identified as a problem but interpreted as a normed response to the contextual experiences (see Swanson, 2010). As such, intervention efforts address the behavior, not the contextual conditionsor the emotional impact (without a primary focus on behavior). Among the sources of trauma for African Americans recently being reintroduced into social science research is cultural trauma. The intergenerational impact of cultural trauma, shaped by the historical experiences discussed earlier, draws on critical race theory and extends challenges regarding inferences based on reactive behavior. According to Stamm, Stamm, Hudnall, and Higson-Smith (2004), cultural trauma is nonspecific and tends to help explain human interchange across cultures with differences in economic, social, and technological organization. Sociocultural trauma exists because of certain conditions either inherited or forced on people by dominating forces. Over the last few decades, cultural trauma has been examined for many groups worldwide. Alexander, Eyerman, Giesen, Smelser, and Sztompka (2004) defined its impact as “members of a collectivity [who] have been subjected to a horrendous event that leaves indelible marks upon their group consciousness, marking their memories forever and changing their future identity in fundamental and irrevocable ways” (p. 1). Eyerman (2001) argued that “[t]here is a difference between trauma as it affects individuals and as a cultural process. As a cultural process trauma is linked to the formation of collective identity and the construction of collective memory” (p. 1). During the last 20 to 25 years, American media has consistently projected an expectation for citizens to be their own individuals, independently creating the life they envision, and take their piece of the “American Dream.” The literature suggests that this type of socialization does not work well with an ethnic group whose highest value lies in relationships between people (Akbar, 1996; Degruy, 2005). Alexander et al. (2004) suggested that an understanding of cultural trauma is relevant in exploring how it contributes to, or affects, everyday life. This framework implies that socialcultural traumas are fused into the collective identity and collective memory via generational transmission influencing the individual development and self-worth of those impacted. Although it is implied in relation to impacting socialization practices and youth outcomes, limited research connects parenting patterns and cultural socialization or practices used to raise children. Training New Scholars The 30 years of research on vulnerability and resilience created a shift from pathologically based studies to greater emphasis on developmental and contextual framing. In essence, earlier research involving African American children and adolescents was based on comparisons to Whites, was atheoretical, lacked developmental considerations, and was based on a priori assumptions regarding deviance (see Spencer 1985, 1990a). A shift toward exploring PYD offered hope in addressing these earlier issues: Studies explore within-group development, are theoretically conceptualized, and began to consider developmental processes. Since a priori

assumptions remained unchallenged, the research continues to replicate biased assumptions. There is significant research on within-group processes for African American youth; a wealth of research also is available using multi-ethnic samples in which African American youth are compared to other youth. Limited developmental understanding is applied to exploring the social-cognitive and culturally based, meaning-making processes necessary to interpret contextual influences impacting youth outcomes. These trends have not only influenced the focus and nature of research but have also shaped the training of emerging scholars who will continue in this “tradition” of conceptualization. The American Psychological Association (2008) task force guidelines for multicultural training and research recommend that consideration is given to the impact of researcher perceptions on implementing a research plan. These include the need to recognize: a. that psychologists are cultural beings that “may hold attitudes and beliefs that can detrimentally influence their perceptions of and interactions with individuals who are ethnically and racially different from themselves”; b. “the importance of multicultural sensitivity/responsiveness, knowledge, and understanding about ethnically and racially different individuals”; and c. the importance of “conducting culture-centered and ethical psychological research among people from ethnic, linguistic, and racial minority backgrounds.” (pp. 3–18) The guidelines are to encourage an awareness of ethical research in conducting studies with diverse groups, but they fail to highlight the need for researchers to examine personal biases and perspectives of the groups being examined from the conception of a study through analysis and interpretation. As such, subsequent emerging scholars will not have an awareness of how positions of privilege and perceptions of others impact their research that can inform policies and practices for youth of color generally and African American youth specifically. Without such insights, researchers themselves are at risk for using publicly funded and federal research dollars to foster successful careers while perpetrating perspectives that compromise the potential effectiveness of policies and programs for youth and families most in needed of authentic supports. As the examples offered suggest, integrating a developmental perspective into research on diverse groups requires an articulated understanding of the processes by which development takes place. In addition to the collective impact of historical experiences and cognitive maturation as relevant, these constructs are seldom explicitly examined in studies of vulnerability and resilience. They are also minimally addressed as an interpretational consideration that is particularly relevant when examining the meaning-making processes involved in negotiating experiences of stress.

Translational Practices Since the mid-1990s, federal and foundation funding agencies have increasingly required accountability for the dissemination of results obtained from supported studies. These agencies

have been linked to policies requiring explicit collaborations between researchers and providers along with a public expectation for research to inform services being provided or obtained. During President Johnson's administration, policies meant to address poverty were framed to provide opportunities for youth and families (see Swanson & Spencer, 1991). These practices, viewed by a later administration as an economic drain, were eliminated. Subsequent policies since the 1990s have focused on addressing families and youth as problems to be fixed or managed, resulting in significant attention on individual characteristics and significantly less on societal practices contributing to the noted problems. Consistent with research patterns historically, the developmental field has responded to these policy and societal shifts without shifting the underlying individualistically oriented paradigms. Support for identifying practices that meet the criteria for generalizability (i.e., wide dissemination) while identifying factors relevant for mitigating individual problems are the focus of recent research on resilience. These practices are identified as evidence-based. Evidence-based practices (EBPs) offer best practices for facilitating positive outcomes for youth in various settings drawn from randomized control studies. They have been particularly relevant in addressing targeted mental and physical health risks, but they also are relevant in identifying effective universal strategies as noted earlier with regard to mentoring practices. Numerous standardized programs exist for addressing universal and targeted youth concerns across various institutional systems (J. L. Frank, Jennings, & Greenberg, 2013; Kusche & Greenberg, 2012). Dissemination and extension of these practices has, however, been limited for small-scale prevention and intervention practices, particularly for those needing to ensure cultural relevance. Community-based organizations and faith-based institutions are two areas that provide significant support for youth in which dissemination of findings has been slow. Research has been relatively effective in addressing individual issues, particularly in school and family contexts. Yet even within this framing, there is limited consideration of the developmental needs and cultural factors influencing sustainability of positive outcomes. Specifically, in providing culturally relevant universal interventions, there is no current mandate to adopt a developmental framework in guiding practice-related decisions; a “lack of knowledge and guiding framework can significantly impair [practitioners'] abilities to address the complexities [youth] experience that may call for tailoring or adapting treatment models” (DeRosa, Amaya-Jackson, & Layne, 2013, p. 199). Given the framework of the dual-axis model of vulnerability suggesting that, at some level, all youth are vulnerable (see Spencer, Harpalani, Cassidy, Jacobs, et al., 2006), programs should address youth needs by facilitating coping strategies and competence formation relevant in early adulthood. DeRosa et al. (2013) proposed that an essential approach is one that cultivates both lifelong learning and critical thinking in the knowledge obtained through emerging research and the utilization of evidencebased, trauma-informed assessments. With this knowledge of emgering science and evidencebased practices, programs are better equipped to critically appraise their approaches as they formulate conceptualizations and treatment recommendations for vulnerable youth. Concurrently, however, there is a need to address factors that contribute to youth vulnerability or that continue to exacerbate the concerns. Framing adolescence as a culture in itself, Nelson and Nelson (2010) suggested that EBPs address the multiple components of adolescents' experiences (e.g., technology, peer relationships, identity processes) in developing and

implementing intervention practices. They contended that “Whether modifying existing evidence-based protocols or designing new interventions to be evaluated, thoughtful attention to adolescent culture will help enhance adolescent engagement and promote improved treatment effectiveness” (p. 310). The increase understanding of assets associated with PYD contributes to explicit programmatic considerations of risks and supports. Allison, Edmonds, Wilson, Pope, and Farrell (2011) further suggested that even when prevention or youth development programs are available, community residents may not know about nor access these resources. Community infrastructure (e.g., availability, quality) is important, but resource awareness is also necessary. If we build it and they do not use it, dissemination efforts that just target organizations may not result in ultimate benefits to youth and communities. Support for the prevention delivery system may need to explicitly address those organizational capacities that also engage and link community residents to prevention opportunities and resources. Intentional linkages at each step may be necessary. Next we present two programs that utilize research on vulnerability and resilience in their design and implementation to mitigate challenges and enhance future trajectories of youth. Each program has a specific focus, but both provide a comprehensive approach that includes addressing racial disparities and societal incongruence through developmentally appropriate practices for African American youth. Numerous other programs are based on best practices from the research. We present these two to highlight how program practices can draw from resilience research and to explicitly illustrate the multilevel focus (i.e., individual developmental needs, contextual conditions) required of programs in which research can ensure knowledge of supportive strategies are available. These provide a novel conceptualization in translating research for practice that addresses youth as in need of support (rather than a problem to fix) the developmental tasks relevant for youth, and the contextual supports necessary for a sustainable positive trajectory.

Work Connections Hillside Work-Scholarship Connection (HWSC) is a youth development program designed to help youth (most likely to leave school) to stay in school, graduate, and be prepared for the transition into work or postsecondary education. With several sites nationwide, HWSC assists youth with acknowledging, and achieving their potential to mitigate chronic, negative contextual influences surrounding them. HWSC also endeavors to change the negative educational trajectories of youth by connecting them with resources to become successful. As a secondary intervention, the focus is on youth already showing signs indicative of negative outcomes. Youth are identified based on having at least two risks associated with school dropout: failing a core subject, multiple school suspensions, overage for grade level, high rates of absenteeism, and a family that meets poverty guidelines (Swanson & Spencer, 2011; see Zimmerman et al., 2003, regarding resilience-compromising factors). HWSC defines resilience as students' success in overcoming aversive barriers, created by poverty, as expressed through completing high school and obtaining some post–high school training. To accomplish this, HWSC serves as a bridge that connects its students to

opportunities and resources often blocked or denied by their impoverished contexts. Recognizing its limits in changing students' aversive circumstances, the program positively impacts students' access to profitable skills and knowledge. HWSC offers the traditional resources associated with successful programs addressing academic vulnerabilities of youth: mentoring and academic supports. HWSC's work in translating research to practice, extends into providing work opportunities for youth (i.e., identity-enhancing opportunities), ongoing cultural and developmental training for the staff (i.e., to address perceptions of privilege, personal expectations, and assumptions about development), and training for partner organizations that offer employment (i.e., to address inherent biases in policies or expectations that compromise youth opportunities to learn and grow). Youth are provided and connected to individualized resources to enhance their educational trajectories and the social-emotional development needed to succeed: long-term mentoring, academic support, life and social skills, employment, and postgraduation planning and support. Each component is designed to guide students toward academic achievement by directly addressing contextual barriers and risks (see Bauermeister, Zimmerman, Barnett, & Caldwell, 2007). Long-term mentoring is a critical component to HWSC's overall functioning, and mission. This service is executed by Youth Advocates, who each serves as a mentor to an average caseload of 30 students. To best support and guide students toward the goal of graduation, Youth Advocates (i.e., mentors) build rapport with not only the student but also relevant stakeholders in the student's life, such as family members, teachers, and employers. Through establishing meaningful connections, the Youth Advocate is able to best identify and address each student's specific needs. As part of the life and social skill building service, HWSC staff members conduct weekly enrichments that follow the evidence-based curriculum included within the Teen Outreach Program (TOP; Allen, Spencer, & Brookins, 1985). During these sessions, students are provided opportunities to evaluate current and future life options. These are growth-promoting experiences that exist outside of potential challenges in their homes or neighboring communities. Since many of the students potentially become first-generation college students, they are also provided postsecondary planning and support (e.g., free SAT/ACT prep courses, Free Application for Federal Student Aid [FAFSA] workshops, resume/cover letter building, and organized college tours) to help them manage unfavorable contextual obstacles. Offering students employment is an aspect of the program that encourages focus and a trajectory toward graduation. The skills prepare youth for the transition into future work roles. HWSC has established meaningful relationships with prominent local employers to support job placements. Students must initially meet the program's pre-employment criteria for “Triple-A Standard” (Academics, Attitude, and Attendance), which hold students accountable for average or better grades on recent report cards, good attitude (as denoted by school/program staff), and school attendance above 85%. Once youth are employed, Youth Advocates assist them with time management skills and in balancing obligations at school, work, and home. Employment opportunities through this program instill students with a greater sense of purpose and drive to

succeed. HWSC students who are employed by an employment partner often experience greater access to opportunities and resources, which expands their social, cultural, and financial capital. Factors identified as enhancing youth resilience that are incorporated into this program include mentoring by adults who have shared experiences similar to the youth participants, academic supports, positive peer relationships, and an opportunity for financial autonomy. These factors, identified as particularly relevant for urban African American youth, have been integrated with TOPS as an EBP, to explicitly address the youth needs. An additional component not explicitly explored in developmental research but acknowledged as significant in this chapter is addressing the potential for biases: addressing stereotyped perceptions of the youth and positions of privilege among those who will hire and supervise them. In complementary roles, the Youth Advocate supervisors maintain regular meetings and provide ongoing training for employee supervisors around unintentional biases and youth areas of strengths to utilize and enhance. This component of the program extends the use of existing research into practices addressing systemic biases and discriminatory practices that impact the nature of feedback and work opportunities youth receive consistent with research on youth employment. There is variability in the extent to which these employer meetings occur. However, while a process for establishing fidelity is needed, the initial piloting provides a novel orientation to programmatically maximizing supports for vulnerable youth. The current programmatic and practice guidelines are being refined and will offer potential for exploring translational approaches in the future.

Restorative Practices and Principles Restorative practices, as previously noted, is an approach to addressing behavior problems in schools where the behavior is understood in social context. “Individuals are recognized as being part of a social web of relations, and building, maintaining and repairing relationships become priorities” (Morrison & Vaandering, 2012, p. 138). Restorative principles are based on the premise that harming others damages interpersonal and communal relationships and that justice is best achieved by seeking to repair relationships through reconciliation. Restorative practices, however, evolved from a philosophy of restorative justice that is essentially reactive in response to a disciplinary issue. However, a restorative approach within the context of education moves beyond reactive responses to wrongdoing to include proactive strategies that address the needs of students and staff members to work together in positive relationships as members of a community (Morrison & Vaandering, 2012; Wachtel, 2012): It is a universal intervention approach where behavior is understood in social context in supporting individual development. The school-to-prison pipeline has recently received substantial acknowledgment in various fields but was identified as an emerging issue warranting immediate attention over two decades ago (Mulligan, 2009). This trajectory is traced to progressively harsher penalties for mild to moderate infractions that became linked to juvenile justice proceedings and is associated with discriminatory treatment of minority and marginalized students in many U.S. schools (Caton, 2012; Evans & Lester, 2012; Molsbee, 2008). In an effort to reverse this

process, restorative practice strengthen relationships responsible for the social context or climate. It involves creating an environment where (1) everyone is viewed and treated as valuable and (2) a process for examining the impact of one's behavior on another, beginning with restorative questions [e.g., what did you want (from the disruptive behavior?)], that encourages individual and collective responsibility consistent with cultural values of mutual cooperation and sharing (Murry et al., 2005. It is also conducive for adolescents who have the capacity to reflect and discuss their needs. Loosely drawn from nonviolence principles, questions are framed, and expected to be asked, in a manner supportive of youth development and their relationship with others. Although restorative practices have been used in schools since the 1990s, little research has been done evaluating the effects of these programs using experimental or quasi-experimental methods (Stinchcomb, Bazemore, & Riestenberg, 2006). One exception to this is Wong, Cheng, Ngan, and Ma's research (2011). These authors conducted a quasi-experimental study of a restorative whole-school approach, which found that intervention group participants, as opposed to the partial intervention group and the comparison group that received no intervention, experienced a significant reduction in bullying, higher empathic attitudes, and higher self-esteem. Other studies, examining components of the restorative practice process, suggest that when restorative practices are implemented, behavioral infractions are reduced, suspensions and expulsions decrease, academic achievement improves, school climate and school safety improve, student engagement and connectedness increase, and absenteeism declines (Drewery & Winslade, 2005; International Institute for Restorative Practices, 2009; Karp & Breslin, 2001; McCluskey et al., 2011; Mirsky & Wachtel, 2007; Rideout, Roland, Salinitri, & Frey, 2010; Stinchcomb et al., 2006). Importantly, school connectedness and engagement has been shown for minority students previously identified as having either struggling or as having discipline concerns. Evaluations of discipline approaches in schools show the benefit of collective responsibility in establishing caring relationships and nonpunitive approaches to discipline. Freiberg and Lapointe (2006) evaluated 40 school-based programs that included evaluations of interventions in schools that specifically sought to reduce behavior problems. Of these 40 programs, 29 were implemented in schools with high minority and low-income student demographics. The authors found several commonalities among the successful programs: The programs moved beyond discipline, emphasizing student learning and self-regulation, not simply procedures for addressing rule infractions. They encouraged “school connectedness” and “caring and trusting relationships” between teachers and students (p. 65). Overall, the programs tried to increase students' positive experience of schooling by building mutually respecting relationships for all school staff and students and by moving away from a reliance on punitive reactions to misbehavior. Restorative practices in education offer an approach that manifests itself in a variety of strategies. These practices take a universal approach and systems-based orientation toward creating and sustaining a climate conducive for positive relationships and academic

engagement. It does, however, require substantive shifts in existing practices and policies that can exceed a district's or school's capacity for implementation and sustainability. In addition, the focus on establishing and restoring relationships is based on the premise of mutual respect. There is an expectation that individual differences are minimized in the process, which has been reported as a concern in some contexts. As respect and the nature of relationships are contextually defined, studies are currently exploring the impact of explicitly incorporating training to ensure personal and historical biases are mitigated in the implementation of the practices. Restorative practices do not negate the benefits of EBPs to understanding effective, theoretically framed strategies, yet substantial work is needed to understand and implement these into translational practices that consider contextual and cultural influences on effectiveness. Restorative practice programs are shaped by available research on resilience relevant to their mission and the desired outcomes for their target populations. They acknowledge developmental needs and seek to explicitly integrate cultural influences in their attention to how relationships influence on youth outcomes. Particularly important is attention to providing accepting and caring relationships consistent with research regarding safe and impactful adults in the lives of youth. To varying degrees, restorative practice programs also address individual biases that contribute to youth trajectories and outcomes of concern. Sociohistorical factors related to untoward conditions are generally unknown and therefore unacknowledged. Efforts to explore individual biases and their bases (often from a position of privilege, regardless of race) are ways to better understand programs and their limitations in facilitating resilience among African American youth. System-wide or comprehensive efforts such as HWSC and restorative practices are challenging to implement with the type of fidelity required by EBPs. These types of programs draw from the research to identify what practices to consider as central and where flexibility might be allowed. Smaller programs that exist within community-based organizations and faith-based institutions often lack the resources needed to create conditions that are inclusive of developmental, cultural, and relational factors in order to retrofit them for their mission and youth population. Collectively, small, community level programs reach substantial numbers of youth, and invest significant time, effort, and resources in meeting youth needs with little knowledge of how and why their efforts should work. Translating research into action should involve mechanisms and networks that can explicitly inform those directly involved in meeting youth needs.

Conclusion: Reframing the Future As developmental science is a data-based field that informs decisions within a data-driven society, there is a prevailing assumption that data and data sources represent reality. A primary proposition of this chapter is an acknowledgment that data alone are not always enough. Part of the challenge for research scholars is not only ensuring that their research accurately represents the populations investigated (and the unique historical experiences of each) but, additionally, disseminating the findings to those who can implement the work in a relevant manner.

Acknowledging one's position of privilege and recognizing the need for accurately understanding others' experiences without automatically pathologizing differences is always a necessary first step in the conduct of science. Without that critical consideration, there is the risk of misrepresenting youth and their families due to preconceived and unchallenged assumptions about the meaning and impact of historical and contemporary experiences. Without question, consequent translational work is more complex than developing a treatment plan for problem behaviors. Sufficient data on culturally informed (and thus, historically embedded), contextually based, and developmentally explored phenomena are needed to support research and the implementation of practices. As important, however, is the necessity of examining the influence of investigator privilege and the “wages” of social science status quo adherence on these processes (e.g., the ease of procuring research funds and the securing of journal publishing, given adherence to “traditional perspectives” in the conduct of social science). Multicultural training should be explicitly integrated into research training as a core component of professional ethics. As an extension of the traditional approach to exploring others, a core component of the training must integrate positions of privilege that shape how others are viewed, evaluated, and supported (i.e., resources provided, treatments recommended). Although multicultural training is a course requirement in clinical, counseling, and education programs, there is substantial variation in the extent to which this training is integrated into other core courses and subsequently extended into practice. In addition, multicultural training is not a course or practice requirement in other areas of psychology; frequently it is addressed only within studies exploring research ethics and some justifications associated with the human subjects review process. These approaches to understanding cultural practices suggest they are sufficiently important to be acknowledged but minimally relevant in clinical and research practices. More disconcerting, however, is the perspective that historical and cultural influences are focused on merely to be politically correct. Such an attitude removes personal responsibility to examine and address individual biases from the equation. The need for cultural awareness grew out of grassroot efforts to impact federal policies and professional standards, such as those by the American Psychological Association. Federal policies have focused on nondiscriminatory practices in an effort to facilitate inclusion (e.g., affirmative action) but negated the sociohistorial factors rooted in the attitudes and behaviors that compromised equitable opportunities. Since the mid-1980s, policies influencing African American youth directly focused on addressing observable problems in contrast to the preceding two decades, which focused on providing opportunities: welfare reform did not provide support for the working poor; zero tolerance as a reform for school-based discipline did not mitigate referral biases or create safe environments conducive for learning. As research often follows trends highlighted in national and educational policies, a focus on problems without consideration for the contributing historical or contextual influences has remained in research. There is considerable research regarding factors influencing resilience for African American youth; there is less acknowledgment of practices that continue to constrain their opportunities. Given what is known from basic and applied research, those within the field are positioned to better facilitate resilience by consciously acknowledging societal and professional culpability,

in this way removing systemic barriers that compromise the opportunities and outcomes for African American youth. Until the noted changes occur, the Einstein quote at the beginning of the chapter still holds sway, and virtually guarantees the misrepresentation of African American youth sources of high vulnerability and resilience.

References Akbar, N. (1996). Breaking the chains of psychological slavery. Tallahassee, FL: Mind Productions. Alarcon, O., Szalacha, L. A., Erkut, S., Fields, J. P., & Coll, C. G. (2000). The color of my skin: A measure to assess children's perceptions of their skin color. Applied Developmental Science, 4(4), 208. Alexander, J. C., Eyerman, R., Giesen, B., Smelser, N. J., & Sztompka, P. (2004). Cultural Trauma and Collective Identity. Berkeley, CA: University of California Press. Allen, W. R., Spencer, M. B., & Brookins, G. K. (1985). Synthesis: Black children keep on growing. In M. B. Spencer, G. K. Brookins, & W. R. Allen (Eds.), Beginnings: The social and affective development of Black children (pp. 301–314). Hillsdale, NJ: Erlbaum. Allison, K. W., Edmonds, T., Wilson, K., Pope, M., & Farrell, A. D. (2011). Connecting youth violence prevention, positive youth development, and community mobilization. American Journal of Community Psychology, 48(1–2), 8–20. Altabe, M. (1998). Ethnicity and body image: quantitative and qualitative analysis. International Journal of Eating Disorders, 23(2), 153–159. American Psychological Association, (2008). Report of the Task Force on the Implementation of the Multicultural Guidelines. Washington, DC: Author. Retrieved from http://www.apa.org/pi/ Amstutz, L. S., & Mullet, J. H. (2005). The little book of restorative discipline for schools: Teaching responsibility; creating caring climates. Intercourse, PA: Good Books. Anthony, E. J. (1974). Introduction: The syndrome of the psychologically vulnerable child. In: E. J. Anthony & C. Koupernik (Eds.), The child in his family: Children at Psychiatric Risk. (Vol. 3, pp. 3–10). New York: Wiley. Anthony, E. J. (1987). Risk, vulnerability, and resilience: An overview. In E. J. Anthony & B. J. Cohler (Eds.), The invulnerable child (pp. 3–48). New York, NY: Guilford Press. Aronson, D. N., & Lustina, M. J., Good, C., Keough, K., Steele, C. M., & Brown, J. (1999). When White men can't co math: Necessary and sufficient factors in stereotype threat. Journal of Experimental Social Psychology, 35, 29–46. Aronson, J., Fried, C., & Good, C. (2002). Reducing the effects of stereotype threat on African

American college students by shaping theories of intelligence. Journal of Experimental Social Psychology, 38(2), 113–125. Arrington, E. G., & Wilson, M. N. (2000). A re-examination of risk and resilience during adolescence: incorporating culture and diversity. Journal of Child & Family Studies, 9(2), 221–230. Averhart, C. J., & Bigler, R. S. (1997). Shades of meaning: Skin tone, racial attitudes, and constructive memory in African American children. Journal of Experimental Child Psychology, 67(3), 363–388. doi: 10.1006/jecp.1997.2413 Bachman, J. G., & O'Malley, P. M. (1984). Black-White differences in self-esteem: Are they affected by response styles? American Journal of Sociology, 90(3), 624–639. Bachman, J. G., O'Malley, P. M., Freedman-Doan, P., Trzesniewski, K. H., & Donnellan, M. B. (2011). Adolescent self-esteem: Differences by race/ethnicity, gender, and age. Self and Identity: the Journal of the International Society for Self and Identity, 10(4), 445–473. Barbarin, O. (2013a). Development of boys of color. American Journal of Orthopsychiatry, 83, 143–144. doi: 10.1111/ajop.12032 Barbarin, O. (2013b). A longitudinal examination of socioemotional learning in African American and Latino boys across the transition from pre-K to kindergarten. American Journal of Orthopsychiatry, 83, 156–164. doi: 10.1111/ajop.12024 Barbarin, O., Iruka, I. U., Harradine, C., Winn, D. C., McKinney, M. K., & Taylor, L. C. (2013). Development of social-emotional competence in boys of color: A cross-sectional cohort analysis from pre-K to second grade. American Journal of Orthopsychiatry, 83, 145– 155. doi: 10.1111/ajop.12023 Bauermeister, J. A., Zimmerman, M. A., Barnett, T. E., & Caldwell, C. H. (2007). Working in high school and adaptation in the transition to young adulthood among African American youth. Journal of Youth and Adolescence, 36(7), 877–890. Benson, J. E. (2006). Exploring the racial identities of Black immigrants in the U.S. Sociological Forum, 21(2), 219–247. Benson, P. L. (1990). The troubled journey. Minneapolis, MN: Search Institute. Benson, P. L., & Scales, P. C. (2009). The definition and preliminary measurement of thriving in adolescence. Journal of Positive Psychology, 4(1), 85–104. doi: 10.1080/17439760802399240 Beteille, H. (1968). Race and descent as social categories in India. In J. H. Franklin (Ed.), Color and race (166–185). Boston, MA: Beacon Press. Blankenship, K. M. (1998). A race, class, and gender analysis of thriving. Journal of Social Issues, 54(2), 393–404. doi: 10.1111/j.1540–4560.1998.tb01226.x

Bodenhorn, H. (2006). Colorism, complexion homogamy, and household wealth: Some historical evidence. American Economic Review, 96(2), 256. Bodenhorn, H., & Ruebeck, C. S. (2005). Colorism and African American wealth: Evidence from the nineteenth-century South. National Bureau of Economic Research, Working paper no. 11732. Bonner, F. A. I., Lewis, C. W., Bowman-Perrott, L., Hill-Jackson, V., & James, M. (2009). Definition, identification, identity, and culture: A unique alchemy impacting the success of gifted African American millennial males in school. Journal for the Education of the Gifted, 33(2), 176–202. Boykin, A. W. (1986). The triple quandary and the schooling of African American Children. In U. Neisser (Ed.), The school achievement of minority children: New perspectives (pp. 57– 92). Mahwah, NJ: Erlbaum. Boykin, A. W. (2000). The talent development model of schooling: Placing students at promise for academic success. Journal of Education for Students Placed at Risk, 5(1&2), 3–25. Braithwaite, J. (2002). Setting standards for restorative justice. British Journal of Criminology, 42(3), 563–577. Brittian, A. S., & Spencer, M. B. (2012). Assessing the relationship between ethnic and religious identity among and between diverse American youth. In A. E. A. Warren, R. M. Lerner, & E. Phelps (Eds.), Thriving and spirituality among youth: Research perspectives and future possibilities (pp. 205–230). Hoboken, NJ: Wiley. Bronfenbrenner, U. (1979). The ecology of human development: Experiments by nature and design. Cambridge, MA: Harvard University Press. doi: 10.1080/00131728109336000 Bronfenbrenner, U. (1992). Ecological systems theory. In U. Bronfenbrenner (Ed.), Making human beings human: Bioecological perspectives on human development (pp. 106–173). Thousand Oaks, CA: Sage. Broussard, C., Mosley-Howard, S., & Roychoudhury, A. (2006). Using youth advocates for mentoring at-risk students in urban settings. Children & Schools, 28(2), 122–127. Brown, C. S., & Bigler, R. S. (2005). Children's perceptions of discrimination: A developmental model. Child Development, 76(3), 533–553. doi: 10.1111/j.14678624.2005.00862.x Burke, M. (2008). Colorism. In W. Darity Jr. (Ed.), International encyclopedia of the social sciences (Vol. 2, pp. 17 – 18). Detroit, MI: Thomson Gale. Caldwell, C. H., Guthrie, B. J, & Jackson, J. S. (2006). Identity development, discrimination, and psychological well-being among African American and Caribbean black adolescents. In A. J. Schulz & L. Mullings (Eds.), Gender, race, class, & health: Intersectional approaches

(pp. 163–191). San Francisco, CA: Jossey-Bass. Caton, M. T., (2012). Black male perspectives on their educational experiences in high school. Urban Education, 47(6), 1055–1085. Carter, D. J. (2007). Why the black kids sit together at the stairs: The role of identity-affirming counter-spaces in a predominantly White high school. Journal of Negro Education, 76(4), 542–554. Carter, R. T., Helms, J. E., & Juby, H. L. (2004). The relationship between racism and racial identity for White Americans: A profile analysis. Journal of Multicultural Counseling and Development, 32(1), 2–17. Carter Andrews, D. J. (2009). The construction of black high-achiever identities in a predominantly white high school. Anthropology and Education Quarterly, 40(3), 297–317. Carver, C. S. (1998). Resilience and thriving: Issues, models, and linkages. Journal of Social Issues, 54(2), 245–266. doi: 10.1111/j.1540–4560.1998.tb01217.x Cedeno, L. A., Elias, M. J., Kelly, S., & Chu, B. C. (2010). School violence, adjustment, and the influence of hope on low-income, African American youth. American Journal of Orthopsychiatry, 80(2), 213–226. Charles, C. A. D. (2011). Skin bleaching and the prestige complexion of sexual attraction. Sexuality & Culture: An Interdisciplinary Quarterly, 15(4), 375–390. doi: 10.1007/s12119– 011–9107–0 Chestang, L. W. (1970). The issue of race in casework practice. New York, NY: Columbia University Press. Clark, K. B., & Clark, M. P. (1939). The development of consciousness of self in Negro preschool children. Archives of Psychology. Washington, DC: Howard University. Clark, K. B., & Clark, M. P. (1940). Skin color as a factor in racial identification of Negro preschool children. Journal of Social Psychology, 2, 159–169. doi: 10.1080/00224545.1940.9918741 Constantine, M. G., & Blackmon, S. M. (2002). Black adolescents' racial socialization experiences: Their relations to home, school, and peer self-esteem. Journal of Black Studies, (3), 322. doi: 10.2307/3180866 Cooley, C. H. (1902). Human nature and the social order. New York, NY: Scribner. Cross, W. E., Jr. (1991). Shades of Black: Diversity in African-American identity. Philadelphia, PA: Temple University Press. Cross, W. E., Jr. (2003). Tracing the historical origins of youth delinquency. Journal of Social Issues & Violence: Myths & Realities about Black Culture, 59(1), 67–82.

Cunningham, M., Swanson, D. P., Spencer, M. B., & Dupree, D. (2003). The association of physical maturation with family hassles among African American adolescent males. Journal of Cultural Diversity and Ethnic Minority Psychology, 9(3), 276–288. Darensbourg, A., & Blake, J. (2013). Predictors of achievement in African American students at risk for academic failure: The roles of achievement values and behavioral engagement. Psychology in the Schools, 50(10), 1044–1059. D'Agostino, J. V., Hedges, L. V., Wong, K. K., & Borman, G. D. (2001). Title I parent involvement programs: Effects on parenting practices and student achievement. In G. D. Borman, S. C. Stringfield, & R. E. Slavin (Eds.), Title I: Compensatory education at the crossroads (pp. 117–136). Mahwah, NJ: Erlbaum. Davis, D. B. (1966). The problem of slavery in Western culture. Ithaca, NY: Cornell University Press. Degruy, J. (2005) Post traumatic slave syndrome: America's legacy of enduring injury and healing. Portland, OR: Uptone Press. DeRosa, R. R., Amaya-Jackson, L., & Layne, C. M. (2013). From rifts to riffs: Evidence-based principles to guide critical thinking about next-generation child trauma treatments and training. Training and Education in Professional Psychology, 7(3), 195–204. Deslander, R., & Bertrand, R. (2005). Motivation of parent involvement in secondary-level schooling. Journal of Educational Research, 98, 164–175. DiBartolo, P. M., & Rendón, M. J. (2012). A critical examination of the construct of perfectionism and its relationship to mental health in Asian and African Americans using a cross-cultural framework. Clinical Psychological Review, 32, 139–152. doi: 10.1016/j.cpr.2011.09.007 Diemer, M. A., Kauffman, A., Koenig, N., Trahan, E., & Hsieh, C. A. (2006). Challenging racism, sexism, and social injustice: Support for urban adolescent's critical consciousness development. Cultural Diversity and Ethnic Minority Psychology, 12(3), 444–460. Drewery, W., & Winslade, J. (2005). Developing restorative practices in schools: Some reflections. New Zealand Journal of Counselling, 26(1), 16. DuBois, D. L., & Karcher, M. J. (2005). Youth mentoring: Theory, Research, and Practice. In D. L. DuBois & M. J. Karcher (Eds.), Handbook of youth mentoring (pp. 2–11). Thousand Oaks, CA: Sage. Dupree, D., Spencer, M. B., & Fegley, S. (2007). Perceived social inequity and responses to conflict among diverse youth of color: The effects of social and physical context on youth behavior and attitudes. In R. K. Silbereisen & R. M. Lerner (Eds.), Approaches to positive youth development (pp. 111–131). Thousand Oaks, CA: Sage.

Dupree, D., Spencer, T. R., & Spencer, M. B. (2015).Challenges to and patterned resiliency among African American youth. In L. C. Theron, M. A. Ungar, & L. Liebenberg (Eds.), Resilience and culture(s): Commonalities and complexities. New York, NY: Springer Eccles, J. S., Wong, C. A., & Peck, S. C. (2006). Ethnicity as a social context for the development of African-American adolescents. Journal of School Psychology, 44, 407–426. Edwards, L. M., & Romero, A. J. (2008). Coping with discrimination among Mexican descent adolescents. Hispanic Journal of Behavioral Sciences, 30(1), 24–39. Elkind, D. (1966). Conceptual orientation shifts in children and adolescents. Child Development, 37(3), 493–498. Elkind, D. (1967). Egocentrism in adolescence. Child Development, 38(4), 1025–1034. Ellison, R. (1944). An American dilemma: A review. Teaching American History. Retrieved from http://teachingamericanhistory.org/library/document/an-american-dilemma-a-review/ Elmore, T. G. (2009). Colorism in the classroom: An exploration of adolescents' skin tone, skin tone preferences, perceptions of skin tone stigma and identity. ProQuest Dissertations and Theses, 223. Retrieved from http://search.proquest.com/docview/304983676? accountid=14657 Erikson, E. H. (1950). Childhood & Society. New York, NY: Norton. Evans, A. B., Banerjee, M., Meyer, R., Aldana, A., Foust, M., & Rowley, S. (2012). Racial socialization as a mechanism for positive development among African American youth. Child Development Perspectives, 6(3), 251–257. Evans, K. R., & Lester, J. N. (2012). Zero tolerance: Moving the conversation forward. Intervention in School and Clinic, 48(2), 108–114. Eyerman, R. (2001). Cultural trauma:Slavery and the formation of African American identity. New York, NY: Cambridge University Press, 2001. Falconer, J. W., & Neville, H. A. (2000). African American college women's body image: An examination of body mass, African self-consciousness, and skin color satisfaction. Psychology of Women Quarterly, 24(3), 236–243. doi: 10.1111/j.1471–6402.2000.tb00205.x Fegley, S. G., Spencer, M. B., Goss, T. N., Harpalani, V., & Charles, N. (2008). Colorism embodied: Skin tone and psychosocial well-being in adolescence. In W. Overton, U. Mueller, & J. Newman (Eds.), Developmental perspectives on embodiment and consciousness (pp. 281–311). Mahwah, NJ: LEA. Ferguson, G. M., & Cramer, P. (2007). Self-esteem among Jamaican children: Exploring the impact of skin color and rural/urban residence. Journal of Applied Developmental Psychology, 28(4), 345–359.

Fisher, C. B., Murray, J. P., Dill, J. R., Hagen, J. W., Hogan, M. J., Lerner, R. M., & Wilcox, B. (1993). The national conference on graduate education in the applications of developmental science across the life span. Journal of Applied Developmental Psychology, 14, 1–10. doi: 10.1016/0193–3973(93)90020-V Fisher, C. B., Hoagwood, K., Boyce, C., Duster, T., Frank, D. A., Grisso, T., Levine, R. J., Macklin, R., Spencer, M. B., Takanishi, R., Trimble, J. E., & Zayas, L. H. (2002). Research ethics for mental health science involving ethnic minority children and youths. American Psychologist, 57(12), 1024 –104. Fordham, S., & Ogbu, J. (1986). Black students' school success coping with the burden of “acting white.” Urban Review, 18(3), 176–206. Fouad, N. A., Grus, C. L., Hatcher, R. L., Kaslow, N. J., Hutchings, P. S., Madson, M. B., Crossman, R. B. (2009). Competency benchmarks: A model for understanding and measuring competency in professional psychology across training programs. Training and Education in Professional Psychology, 3(4), S5–S26. Frank, J. L., Jennings, P. A., & Greenberg, M. T. (2013). Mindfulness-based interventions in school settings. Research in Human Development, 10(3), 205–210. Frank, R., Akresh, I. R., & Lu, B. (2010). Latino immigrants and the U.S. racial order: How and where do they fit in? American Sociological Review, 75(3), 378–401. doi: 10.2307/27801532 Freiberg, H. J., & Lapointe, J. M. (2006). Research-based programs for preventing and solving discipline problems. In C. Evertson & C. S. Weinstein (Eds.), Handbook of classroom management: Research, practice, and contemporary issues (pp. 735–786). Mahwah, NJ: Erlbaum. García Coll, C. T., & Magnuson, K. (1999). Cultural influences on child development: Are we ready for a paradigm shift? In A. Masten (Ed.), Minnesota Symposium on Child Psychology (Vol. 29). Mahwah, NJ: Erlbaum. Gaylord-Harden, N. K., Ragsdale, B. L., Mandara, J., Richards, M. H., & Petersen, A. C. (2007). Perceived support and internalizing symptoms in African American adolescents: Selfesteem and ethnic identity as mediators. Journal of Youth & Adolescence, 36(1), 77–88. doi: 10.1007/s10964–006–9115–9 Gibbs, J. T. (1988). Young, Black, and male in America: An endangered species. Dover, MA: Auburn House. Glenn, E. N., Ed. (2009). Shades of difference: Why skin color matters. Stanford, CA: Stanford University Press. Goldsmith, A. H., Hamilton, D., & Darity, W. (2006). Shades of discrimination: Skin tone and wages. American Economic Review, 96 (2), 242–245. doi: 10.2307/30034650

Gonzalez, L. M., Stein, G. L., Kiang, L., & Cupito, A. M. (2014). The impact of discrimination and support on developmental competencies in Latino adolescents. Journal of Latina/o Psychology, 2(2), 79–91. Graham, A., & Anderson, K. A. (2008). “I have to be three steps ahead”: Academically gifted African American male students in an urban high school on the tension between an ethnic and academic identity. Urban Review, 40(5), 472–499. doi: 10.1007/s11256–008–0088–8 Graham, S. (1992). “Most of the subjects were White and middle class”: Trends in published research on African Americans in selected APA journals, 1970–1989. American Psychologist, 47(5), 629–639. doi: 10.1037/0003–066X.47.5.629 Guthrie, R. V. (1976). Even the rat was white: A historical view of psychology. New York, NY: Harper & Row. Gutman, L. M., & McLoyd, V. C. (2000). Parent's management of their children's education within the home, at school and in the community: An examination of African-American families living in poverty. Urban Review, 32, 1–24. Hall, R. E., & Julian Samora Research Institute. (1987). The psychogenesis of color based racism: Implications of projection for dark-skinned Puertorriqueños. East Lansing, MI: Julian Samora Research Institute, Michigan State University. Hanlon, T. E., Simon, B. D., O'Grady, K. E., Carswell, S. B., & Callaman, J. M. (2009). The effectiveness of an after-school program targeting urban African American youth. Education and Urban Society, 42(1), 96–118. Hannon, L., DeFina, R., & Bruch, S. (2013). The relationship between skin tone and school suspension for African Americans. Race and Social Problems, 5(4), 281–295. doi: 10.1007/s12552–013–9104-z Hare, B. R. (1985). Stability and change in self-perception and achievement among Black adolescents: A longitudinal study. Journal of Black Psychology, 11, 29–42. Harpalani, V., Qadafi, A. K., & Spencer, M. B. (2013). Doll studies. In P. L. Mason (Ed.), Encyclopedia of Race and Racism, 2nd ed. (pp. 67–70). Detroit, MI: Cengage. Harpalani, V., & Spencer, M. B. (2009). Status. In R. A. Shweder (Ed.), The child: An encyclopedic companion (pp. 954–956). Chicago, IL: University of Chicago Press. Harper, S. R. (2006). Peer support for African American male college achievement: Beyond internalized racism and the burden of “acting White.” Journal of Men's Studies, 14(3), 337– 358. Harris, C. (1995). Whiteness as property. In K. Crenshaw, N. Gotanda, G. Peller, & K. Thomas (Eds.), Critical race theory: The key writings that formed the movement (pp. 276– 291). New York, NY: New Press.

Harris, C. I. (2002). Critical race studies: An introduction. UCLA Law Review, 49(5), 1215– 1239. Harris-Britt, A., Valrie, C. R., Kurtz-Costes, B., & Rowley, S. J. (2007). Perceived racial discrimination and self-esteem in African American youth: Racial socialization as a protective factor. Journal of Research on Adolescence, 17(4), 669–682. doi: 10.1111/j.1532– 7795.2007.00540.x Harrison, M. S., & Thomas, K. M. (2009). The hidden prejudice in selection: A research investigation on skin color bias. Journal of Applied Social Psychology, 39(1), 134–168. doi: 10.1111/j.1559–1816.2008.00433.x Hatcher, R. L., Fouad, N. A., Grus, C. L., Campbell, L. F., McCutcheon, S. R., & Leahy, K. L. (2013). Competency benchmarks: Practical steps toward a culture of competence. Training and Education in Professional Psychology, 7(2), 84–91. Hatchett, B. F., Holmes, K., Duran, D. A., & Davis, C. (2000). African Americans and research participation: The recruitment process. Journal of Black Studies, 30(5), 664–675. Havighurst, R. J. (1953). Human development and education. New York, NY: McKay. Helms, J. E. (Ed.) (1990). Black and white racial identity: Theory, research, and practice. New York, NY: Greenwood Press, 1990. Herrera, C., Grossman, J. B., Kauh, T. J., Feldman, A. F., & McMaken, J. (2007). Making a difference in schools: The Big Brothers Big Sisters school-based mentoring impact study. Philadelphia: Public/Private Ventures. Hersch, J. (2011). The persistence of skin color discrimination for immigrants. Social Science Research, 40(5), 1337–1349. doi: 10.1016/j.ssresearch.2010.12.006 Hill, S. A. (2001). Class, race, and gender dimensions of child rearing in African American families. Journal of Black Studies, 31, 494–508. Hill, M. (2002). Skin color and the perception of attractiveness among African Americans: Does gender make a difference? Social Psychology Quarterly, 65(1), 77–91. Horowitz, R. E. (1939). Racial aspects of self-identification in nursery school children. Journal of Psychology, 7, 91–99. Hughes, D., & Johnson, D. (2001). Correlates in children's experiences of parents' racial socialization behaviors. Journal of Marriage and Family, 63(4), 981. Hughes, M., & Demo, H. (1989). Self-perception of Black Americans: Self-esteem and personal efficacy. American Journal of Sociology, 95, 132–159. Hughes, M., & Hertel, B. R. (1990). The significance of color remains: A study of life chances, mate selection, and ethnic consciousness among Black Americans. Social Forces, 68(4),

1105–1120. doi: 10.2307/2579136 Hunn, V. (2014). African American students, retention, and team-based learning: A review of the literature and recommendations for retention at predominately White institutions. Journal of Black Studies, 45(4), 301–314. Hunter, M. L. (2002). “If you're light you're alright”: Light skin color as social capital for women of color. Gender and Society, 16(2), 175. International Institute for Restorative Practices. (2009). Improving school climate: Findings from schools implementing restorative practices. Retrieved July 22, 2009, from: http://www.iirp.org/pdf/IIRP-Improving-School-Climate.pdf James, S. A., Keenan, N. L., Strogatz, D. S., Browning S. R., & Garrett, J. M. (1992). Socioeconomic status, John Henryism, and blood pressure in Black adults: The Pitt County Study. American Journal of Epidemiology, 135, 59–67. Jordan, P., & Hernandez-Reif, M. (2009). Reexamination of young children's racial attitudes and skin tone preferences. Journal of Black Psychology, 35(3), 388–403. doi: 10.1177/0095798409333621 Jung, T., & Wickrama, K. A. S. (2008). An introduction to latent class growth analysis and growth mixture modeling. Social and Personality Psychology Compass, 2, 302–317. Karp, D. R., & Breslin, B. (2001). Restorative justice in school communities. Youth & Society, 33(2), 249–272. doi: 10.1177/0044118X01033002006 Kaufman, E. A., & Wiese, D. L. (2012). Skin-tone preferences and self-representation in Hispanic children. Early Child Development and Care, 182(2), 277–290. doi: 10.1080/03004430.2011.556250 Keith, V. M., & Herring, C. (1991). Skin tone and stratification in the Black community. American Journal of Sociology, 97(3), 760–778. doi: 10.1086/229819 Kiang, L., & Takeuchi, D. T. (2009). phenotypic bias and ethnic identity in Filipino Americans. Social Science Quarterly, 90(2), 428–445. doi: 10.1111/j.1540–6237.2009.00625.x Kinzler, K. D., & Dautel, J. B. (2012). Children's essentialist reasoning about language and race. Developmental Science, 15(1), 131–138. Kinzler, K. D., & Shutts, K. (2008). Memory for “mean” over “nice”: The influence of threat on children's face memory. Cognition, 107(2), 775–783. Kinzler, K. D., Shutts, K., & Correll, J. (2010). Priorities in social categories. European Journal of Social Psychology, 40(4), 581–592. Kohn, A. (2006). Beyond discipline: From compliance to community. Alexandria, VA: Association for Supervision and Curriculum Development.

Kusche, C. A., & Greenberg, M. T. (2012). The PATHS Curriculum: Promoting emotional literacy, prosocial behavior, and caring classrooms. In S. R. Jimerson, A. B., Nickerson, M. J. Mayer, & M. J. Furlong (Eds.), Handbook of school violence and school safety: International research and practice (2nd ed.) (pp. 435–446). New York, NY: Routledge/Taylor & Francis. Lambert, S., Herman, K., Bynum, M., & Ialongo, N. (2009). Perceptions of racism and depressive symptoms in African American adolescents: The role of perceived academic and social control. Journal of Youth and Adolescence, 38(4), 519–531. doi: 10.1007/s10964– 009–9393–0 Landor, A. M., Simons, L. G., Simons, R. L., Brody, G. H., Bryant, C. M., Gibbons, F. X.,… Melby, J. N. (2013). Exploring the impact of skin tone on family dynamics and race-related outcomes. Journal of Family Psychology, 27(5), 817–826. doi: 10.1037/a0033883 Lee, C. D., Spencer, M. B., & Harpalani, V. (2003). “Every shut eye ain't sleep”: Studying how people live culturally. Educational Researcher, 32(5), 6–13. Lerner, R. M., Lerner, J. V., von Eye, A., Bowers, E. P., & Lewin-Bizan, S. (2011). Individual and contextual bases of thriving in adolescence: A view of the issues. Journal of Adolescence, 34(6), 1107–1114. doi: 10.1016/j.adolescence.2011.08.001 Lewis, C., Butler, B., Bonner, F., & Joubert, M. (2010). African American male discipline patterns and school district responses resulting impact on academic achievement: Implications for urban educators and policy makers. Journal of African American Males in Education, 1(1), 8–25. Lindsay, C. A. (2011). All middle-class families are not created equal: Explaining the contexts that Black and White families face and the implications for adolescent achievement. Social Science Quarterly, 92(3), 761–781. López, I. (2008). “But you don't look Puerto Rican”: The moderating effect of ethnic identity on the relation between skin color and self-esteem among Puerto Rican women. Cultural Diversity and Ethnic Minority Psychology, 14(2), 102–108. doi: 10.1037/1099– 9809.14.2.102 Lusk, E. M., Taylor, M. J., Nanney, J. T., & Austin, C. C. (2010). Biracial identity and its relation to self-esteem and depression in mixed Black/White biracial individuals. Journal of Ethnic and Cultural Diversity in Social Work, 19(2), 109–126. doi: 10.1080/15313201003771783 Luthar, S. S., & Lattendresse, S. J. (2002). Adolescent risk: The cost of affluence. New Directions for Youth Development, 95, 101–121. doi: 10.1002/yd.18 MacLeod, J. (1995). Ain't no makin' it: aspirations and attainment in a low-income neighborhood. Boulder, CO: Westview Press.

Marcia, J. (1966). Development and validation of ego-identity status. Journal of Personality and Social Psychology, 3(5), 551–558. Marcia, J. (1980). Identity in adolescence. In J. Anderson (ed.), Handbook of adolescent psychology. New York, NY: Wiley & Sons. (pp. 159–187). Marshall, S. (1995). Ethnic socialization of African American children: Implications for parenting, identity development, and academic achievement. Journal of Youth and Adolescence, 24, 377–396. Masten, A. S., Hubbard, J. J., Gest, S. D., Tellegen, A., Garmezy, N., & Ramireza, M. (1999). Competence in the context of adversity: Pathways to resilience and maladaptation from childhood to late adolescence. Development and Psychopathology, 11, 143–169 McArdle, C. G., & Young, N. E. (1970). Clashing description of racial identity or how can we make it without “acting White.” American Journal of Orthopsychiatry, 40(1), 135–144. McCluskey, G., Kane, J., Lloyd, G., Stead, J., Riddell, S., & Weedon, E. (2011). “Teachers are afraid we are stealing their strength”: A risk society and restorative approaches in school. British Journal of Educational Studies, (2), 105. doi: 10.2307/41287883 McCluskey, G., Stead, J., Kane, J., Riddell, S., & Weedon, E. (2008). “I was dead restorative today”: From restorative justice to restorative approaches in school. Cambridge Journal of Education, 38(2), 199–216. doi: 10.1080/03057640802063262 McDonald, J. A. (2006). Potential influence of racism and skin tone on early personality formation. Psychoanalytic Review, 93(1), 93–116. doi: 10.1521/prev.2006.93.1.93 McGee, E. O., & Spencer, M. B. (2013). “Going from all Black to predominantly White”: Black students raised in Black urban neighborhoods that attend PWIs. In C. C. Yeakey, V. S. Thompson, & A. Wells (Eds.), Urban ills: Post Recession complexities of urban living in the twenty first century. Lanham, MD: Lexington Books. McGee, E. O., Hall, J., & Spencer, M. B. (2015). Black parents as advocates, motivators, and teachers of mathematics. In N. Nasir, C. Lee, R. Pea, et al. (Eds.), The handbook for the cultural foundations of learning. McHale, S. M., Crouter, A. C., Kim, J.-Y., Burton, L. M., Davis, K. D., Dotterer, A. M., & Swanson, D. P. (2006). Mothers' and fathers' racial socialization in African American Families: Implications for youth. Child Development, 77(5), 1387–1402. McIntosh, P. (July–August 1989). White privilege: Unpacking the invisible knapsack. Peace and Freedom, 9–10. McNeil, S., Harris-McKoy, D., Brantley, C., Fincham, F., & Beach, S. R. (2014). Middle class African American mothers' depressive symptoms mediate perceived discrimination and reported child externalizing behaviors. Journal of Child and Family Studies, 23(2), 381–388.

Mello, Z. R., & Swanson, D. P. (2007). Gender differences in African American adolescents' personal, educational, and occupational expectations and perceptions of neighborhood quality. Journal of Black Psychology, 33(2), 150–168. Mirsky, L., & Wachtel, T. (2007). The worst school i've ever been to: Empirical evaluations of a restorative school and treatment milieu. Reclaiming Children & Youth, 16(2), 13–16. Molsbee, S. (2008). Zeroing out zero tolerance: Eliminating zero tolerance policies in Texas schools. Texas Tech Law Review, 40, 325–363. Monroe, C. R. (2013). Colorizing educational research: African American life and schooling as an exemplar. Educational Researcher, 42(1), 9–19. doi: 10.3102/0013189X12469998 Montalvo, F. (2004). Surviving race: Skin color and the socialization and acculturation of Latinas. Journal of Ethnic & Cultural Diversity in Social Work, 13(3), 25. Morrissey, K. M., & Werner-Wilson, R. J. (2005). The relationship between out-of-school activities and positive youth development: An investigation of the influences of communities and family. Adolescence, 40(157), 67–85. Morrison, B., E., & Vaandering, D. (2012). Restorative justice: Pedagogy, praxis, and discipline. Journal of School Violence, 11(2), 138–155. doi: 10.1080/15388220.2011.653322 Mulligan, S. (2009). From retribution. La Verne Law Review, 31(1), 139. Murry, V. M., Berkel, C., Brody, G. H., Miller, S. J., & Chen, Y. (2009). Linking parental socialization to interpersonal protective processes, academic self-presentation, and expectations among rural African American youth. Cultural Diversity and Ethnic Minority Psychology, 15(1), 1–10. doi: 10.1037/a0013180 Murry, V. M., Brody, G. H., McNair, L. D., Luo, Z., Gibbons, F. X., Gerrard, M., et al. (2005). Parental involvement promotes rural African American youths' self-pride and sexual selfconcepts. Journal of Marriage & Family, 67, 627–642. Murry, V. M., Smith, E. P., & Hill, N. E. (2001). Race, ethnicity, and culture in studies of families in context. Journal of Marriage & Family, 63(4), 911. Muuss, R. E., Porton, H., & Velder, E. (1996). Theories of adolescence. New York, NY: McGraw-Hill. Myrdal, G. (1944). An American dilemma: The Negro problem and modern democracy. New York, NY: Harper & Row. Nakamura, N., Tummala-Narra, P., & Zárate, M. A. (2013). Expanding our borders: Cultural Diversity and Ethnic Minority Psychology's special issue on immigration. Cultural Diversity and Ethnic Minority Psychology, 19(3), 233–235. doi: 10.1037/a0032959 Neal, D., & Rick, A. (2014, February 15). The prison boom & the lack of black progress

after Smith & Welch. Working paper. Nelson, T. D., & Nelson, J. M. (2010). Evidence-based practice and the culture of adolescence. Professional Psychology: Research and Practice, 41(4), 305–311. Noguera, P. (2008). The trouble with black boys and other reflections on race, equity, and the future of public education. San Francisco, CA: Wiley & Sons. O'Connor, C., Hill, L. D., & Robinson, S. R. (2009). Who's at risk in school and what's race got to do with it? Review of Research in Education, 33(1),1–34. Ogbu, J. U. (1978). Minority education and caste: The American system in cross-cultural perspective. New York: Academic Press. Ogbu, J. U., & Simons, H. D. (1998). Voluntary and involuntary minorities: A culturalecological theory of school performance with some implications for education. Anthropology & Education Quarterly 29(2), 155–188. O'Leary, V. E., & Ickovics, J. R. (1995). Resilience and thriving in response to challenge: an opportunity for a paradigm shift in women's health. Women's Health, 1(2), 121–142. Oney, C., Cole, E., & Sellers, R. (2011). Racial identity and gender as moderators of the relationship between body image and self-esteem for African Americans. Sex Roles, 65(7/8), 619–631. doi: 10.1007/s11199–011–9962-z Parsons, C. (2005). School exclusion: The will to punish. British Journal of Educational Studies, 53(2), 187–211. Phelan, P., Davidson, A. L., & Cao, H. T. (1991). Students' multiple worlds: Negotiating the boundaries of family, peer, and school cultures. Anthropology & Education Quarterly, 22, 224–250. Phinney, J. S., Cantu, C. L., & Kurtz, D. (1997). Ethnic and American identity as predictors of self-esteem among African American, Latino, and White adolescents, 26(2), 165–185. doi: 10.1023/A:1024500514834 Pina-Watson, B., Ojeda, L., Castellon, N. E., & Dornhecker, M. (2013). Familismo, ethnic identity, and bicultural stress as predictors of Mexican American adolescents' positive psychological functioning. Journal of Latina/o Psychology, 1(4), 204–217. Pollard, D. S. (1989). Against the odds: A Profile of academic achievers from the urban underclass. Journal of Negro Education, 58(3), 297–308. Popa, C. N. (2012). Restorative justice: A critical analysis. Law Review: Judicial Doctrine & Case-Law, 2(3), 2. Porter, J. R., & Washington, R. E. (1989). Developments in research on Black identity and self-esteem: 1979–1988. Revue Internationale de Psychologie Sociale, 2(3), 339–353.

Prilleltensky, I. (2008). The role of power in wellness, oppression, and liberation: The promise of psychopolitical validity. Journal of Community Psychology, 36(2), 116–136. doi: 10.1002/jcop.20225 Quiros, L., & Dawson, B. A. (2013). The color paradigm: The impact of colorism on the racial identity and identification of Latinas. Journal of Human Behavior in the Social Environment, 23(3), 287–297. doi: 10.1080/10911359.2012.740342 Resnicow, K., Soler, R. E., Braithwaite, R. L., Selassie, M. B., & Smith, M. (1999). Development of a racial and ethnic identity scale for African American adolescents: The survey of Black life. Journal of Black Psychology, 25(2), 171–188. doi: 10.1177/0095798499025002003 Rideout, G., Roland, K., Salinitri, G., & Frey, M. (2010). Measuring the impact of restorative justice practices: Outcomes and context. Journal of Educational Administration and Foundations, 21, 35–60. Rockquemore, K. A., & Brunsma, D. L. (2002). Socially embedded identities: Theories, typologies, and processes of racial identity among black/white biracials. Sociological Quarterly, 43, 335–356. Roediger, D. R. (1991). The wages of whiteness: Race and the making of the American working class. New York, NY: Verso. Rogers, M. R. (2006). Exemplary multicultural training in school psychology programs. Cultural Diversity and Ethnic Minority Psychology, 12(1), 115–133. Rowley, S. J., Sellers, R. M., Chavous, T. M., & Smith, M. A. (1998). The relationship between racial identity and self-esteem in African American college and high school students. Journal of Personality and Social Psychology, 74(3), 715–724. Russell, K., Wilson, M., & Hall, R. E. (1993). The color complex: the politics of skin color among African Americans. New York, NY: Anchor Books. Rutter, M. (1987). Psychosocial resilience and protective mechanisms. American Journal of Orthopsychiatry, 57, 316–331. Ryabov, I. (2013). Colorism and school-to-work and school-to-college transitions of African American adolescents. Race and Social Problems, 5(1), 15–27. doi: 10.1007/s12552–012– 9081–7 Sahay, S., & Piran, N. (1997). Skin-color preferences and body satisfaction among South Asian-Canadian and European-Canadian female university students. Journal of Social Psychology, 137, 161–171. doi: 10.1080/00224549709595427 Scales, P. C., Benson, P. L., Leffert, N., & Blyth, D. A. (2000). The contribution of developmental assets to the prediction of thriving among adolescents. Applied Developmental

Science, 4, 27–46. Scales, P. C., Benson, P. L., & Roehlkepartain, E. C. (2011). Adolescent thriving: The role of sparks, relationships, and empowerment. Journal of Youth and Adolescence, 40(3), 263–277. doi: 10.1007/s10964–010–9578–6 Schwartz, S. J., Syed, M., Yip, T., Knight, G. P., Umana-Taylor, A. J., Rivas-Drake, D., & Lee, R. M. (2014). Methodological issues in ethnic and racial identity research with ethnic minority populations: Theoretical precision, measurement issues, and research designs. Child Development, 85(1), 58–76. Seaton, E. K., Caldwell, C. H., Sellers, R. M., & Jackson, J. S. (2008). The prevalence of perceived discrimination among African American and Caribbean Black youth. Developmental Psychology, 44(5), 1288–1297. Sellers, R. M., Smith, M. A., Shelton, J. N., Rowley, S. A. J., & Chavous, T. M. (1998). Multidimensional model of racial identity: A reconceptualization of African American racial identity. Personality & Social Psychology Review, 2(1), 18. Siddle-Walker, V. (1996). Their highest potential: An African American school community in the segregated south. Chapel Hill, NC: University of North Carolina Press. Simmons, R. G., Brown, L., Bush, D. M., & Blyth, D. A. (1978). Self-esteem and achievement of Black and White adolescents. Social Problems, 26(1), 86–96. Small, M. L., Harding, D. J., & Lamont, M. (2010). Reconsidering culture and poverty. Annals of the American Academy of Political and Social Science, 629, 6–27. Sohoni, D., & Saporito, S. (2009). Mapping school segregation: Using GIS to explore racial segregation between schools and their corresponding attendance areas. American Journal of Education, 115(4), 569–600. Solorzano, D., Ceja, M., & Yosso, T. (2000). Critical race theory, racial microagressions, and campus racial climate: The experiences of African American college students. Journal of Negro Education, 69, 60–73. Spanierman, L. B., Poteat, V. P., Beer, A. M., & Armstrong, P. I. (2006). Psychosocial costs of racism to Whites: Exploring patterns through cluster analysis. Journal of Counseling Psychology, 53(4), 434–441. Spencer, M. B. (1976). The competence model as a viable alternative to IQ test gamesmanship. Proceedings of the Association of Black Psychologists. Chicago. Spencer, M. B. (1982a). Personal and group identity of Black children: An alternative synthesis. Genetic Psychology Monographs, 106(1), 59–84. Spencer, M. B. (1982b). Preschool children's social cognition and cultural cognition: A cognitive developmental interpretation of race dissonance findings. Journal of Psychology,

112(2), 275–286. Spencer, M. B. (1985). Cultural cognition and social cognition as identity factors in Black children's personal-social growth. In M. B. Spencer, G. K. Brookins, & W. R. Allen (Eds.), Beginnings: The social and affective development of Black children (pp. 215–230). Hillsdale, NJ: Erlbaum. Spencer, M. B. (1990a). Development of minority children: An introduction. Child Development, 61(2), 267–269. Hillsdale, NJ: Erlbaum. Spencer, M. B. (1990b). Parental values transmission: Implications for the development of African-American children. In H. Cheathan & J. B. Stewart (Eds.), Black families: Interdisciplinary perspectives (pp. 111–130). Atlanta, GA: Transactions. Spencer, M. B. (1995). Old issues and new theorizing about African American youth: A phenomenological variant of ecological systems theory. In R. L. Taylor (Ed.), AfricanAmerican youth: Their social and economic status in the United States (pp. 37–69). Westport, CT: Praeger. Spencer, M. B. (2006). Phenomenology and ecological systems theory: Development of diverse groups. In R. M. Lerner & W. Damon (Eds.), Handbook of child psychology, Vol. 1: Theoretical models of human development, 6th ed. (pp. 829–893). Hoboken, NJ: Wiley. Spencer, M. B. (2008a). Phenomenology and ecological systems theory: Development of diverse groups. In W. Damon & R. M. Lerner (Eds.), Child and adolescent development: An advanced course (pp. 696–735). Hoboken, NJ: Wiley. Spencer, M. B. (2008b). Lessons learned and opportunities ignored since Brown v. Board of Education: Youth development and the myth of a color-blind society (Fourth annual Brown lecture in education research). Educational Researcher, 37(5), 253–266. Spencer, M. B., Brookins, G. K., & Allen, W. R. (Eds) (1985). Beginnings: Social and affective development of Black children. Hillsdale, NJ: Erlbaum. Spencer, M. B., Cole, S. P., Dupree, D., Glymph, A., & Pierre, P. (1993). Self-efficacy among urban African American early adolescents: Exploring issues of risk, vulnerability and resilience. Development and Psychopathology, 5, 719–739. Spencer, M. B., Cross, W. E., Harpalani, V., & Goss, T. N. (2003). Historical and developmental perspectives on Black academic achievement: Debunking the “acting White” myth and posing new directions for research. In C. C. Yeakey & R. D. Henderson (Eds.), Surmounting all odds: Education, opportunity and society in the new millennium (pp. 273– 304). Greenwich, CT: Information Age. Spencer, M. B., & Dornbusch, S. (1990). Challenges in studying minority youth. In S. Feldman & G. Elliot (Eds.), At the threshold: The developing adolescent (pp. 123–146). Cambridge, MA: Harvard University Press.

Spencer, M. B., Dupree, D., Cunningham, M., Harpalani, V., & Muñoz-Miller, M. (2003). Vulnerability to violence: A contextually-sensitive, developmental perspective on African American adolescents. Journal of Social Issues, 59(1), 33–49. Spencer, M. B., Fegley, S., and Harpalani, V. (2003). A theoretical and empirical examination of identity as coping: Linking coping resources to the self processes of African American youth. Journal of Applied Developmental Science, 7(3), 181–187. Spencer, M. B., & Harpalani, V. (2008). What does “acting White” really mean? In J. Ogbu (Ed.), Minority status, oppositional culture, and status (pp. 223–239). New York: Routledge. Spencer, M. B., Harpalani, V., Cassidy, E., Jacobs, C., Donde, S., Goss, T. N.,… Wilson, S. (2006). Understanding vulnerability and resilience from a normative development perspective: Implications for racially and ethnically diverse youth. In D. Cicchetti & D. J. Cohen (Eds.), Handbook of developmental psychopathology, Vol. 1: Theory and method, 2nd ed. (pp. 627– 672). Hoboken, NJ: Wiley. Spencer, M. B., & Horowitz, F. D. (1973). Effects of systematic social and token reinforcement on the modification of racial and color concept attitudes in black and in white preschool children. Developmental Psychology, 9, 246–254. Spencer, M. B., & Markstrom-Adams, C. (1990). Identity processes among racial and ethnic minority children in America. Child Development, 61(2), 290–310. Spencer, M. B., Noll, E., Stoltzfus, J., & Harpalani, V. (2001). Identity and school adjustment: Revisiting the “acting White” assumption. Educational Psychologist, 36(1), 21–30. Spencer, M. B., & Spencer, T. R. (2015). Commentary: Exploring the promises, intricacies, and challenges to positive youth development. Journal of Youth and Adolescence, 43(6), 1027–1033. Spencer, M. B., & Swanson, D. P. (2013). Opportunities and challenges to the development of healthy children living in diverse communities, Developmental and Psychopathology, 25, 1551–1566. Spencer, M. B., Swanson, D. P., & Harpalani, V. (2015). Conceptualizing the self: Contributions of normative human processes, diverse contexts and social opportunity. In M. Lamb, C. G. Coll, & R. Lerner (Eds.), Handbook of child psychology and developmental science (pp. 750–793). Hoboken, NJ: Wiley. Spencer, M. B. and Tinsley, B. (2008). Identity as coping: Youths' diverse strategies for successful adaptation, Prevention Researcher, 15(4), 17–21. Stamm, B. H., Stamm, H. E., Hudnall, A. C., & Higson-Smith, C. (2004). Considering a theory of cultural trauma and loss. Journal of Loss & Trauma, 9(1), 89–111. doi: 10.1080/15325020490255412

Steele, C. M. (1997). A threat in the air. American Psychologist, 52(6), 613–629. Stepanova, E. V., & Strube, M. J. (2009). Making of a face: Role of facial physiognomy, skin tone, and color presentation mode in evaluations of racial typicality. Journal of Social Psychology, 149(1), 66–81. doi: 10.3200/SOCP.149.1.66–81. Stevenson, H. C. (1997). “Missed, dissed, and pissed:” Making meaning of neighborhood risk, fear and anger management in urban black youth. Cultural Diversity and Mental Health, 3(1), 37–52. Stevenson, H. C., & Arrington, E. G. (2009). Racial/ethnic socialization mediates perceived racism and the racial identity of African American adolescents. Cultural Diversity and Ethnic Minority Psychology, 15(2), 125–136. Stinchcomb, J., Bazemore, G., & Riestenberg, N. (2006). Beyond zero tolerance: Restoring justice in secondary schools. Youth Violence & Juvenile Justice, 4(2), 123–147. Stokes-Guinan, K. (2011). Age and skin tone as predictors of positive and negative racial attitudes in Hispanic children. Hispanic Journal of Behavioral Sciences, 33(1), 3–21. doi: 10.1177/0739986310389303 Swain, A., & Noblit, G. (2011). Education in a punitive society: An introduction. Urban Review, 43(4), 465–475. doi: 10.1007/s11256–011–0186-x Swanson, D. P. (2010). Psychosocial development: Identity, stress and competence. In D. P. Swanson, M. C. Edwards, & M. B. Spencer (Eds.), Adolescence: Development in a global era (pp. 93–121). Boston, MA: Elsevier. Swanson, D. P., & Spencer, M. B. (2011). Competence formation: Framing resilience for adolescents' academic outcomes. In K. S. Gallagher, D. Brewer, R. Goodyear, & E. Bensimon (Eds.), Introduction to Urban Education. London, UK: Routledge. Swanson, D. P., Cunningham, M., & Spencer, M. B., (2003). Black males' structural conditions, achievement patterns, normative needs, and “opportunities.” Urban Education Journal, 38(5), 608–633. Swanson, D. P., Cunningham, M., Youngblood, J., & Spencer, M. B. (2009). Racial identity development during childhood. In H. Neville, B. Tynes, & S. O. Utsey (Eds.), Handbook of African American psychology (pp. 269–281). Thousand Oaks, CA: Sage. Swanson, D. P., & Spencer, M. B. (1991). Youth policy, poverty, and African-Americans: Implications for resilience. Education and Urban Society, 24(1), 148–161. Swanson, D. P., & Spencer, M. B. (1999). Developmental and cultural context considerations for research on African American adolescents. In H. Fitzgerald, B. M. Lester, & B. Zuckerman (Eds.), Children of color: Research, health and public policy issues (pp. 53–72). New York, NY: Garland.

Swanson, D. P., Spencer, M. B., Dell'Angleo, T., Harpalani, V., & Spencer, T. R. (2002). Identity processes and the positive youth development of African Americans: An explanatory framework. In C. S. Taylor, R. M. Lerner, & A. von Eye (Eds.), Pathways to positive youth development among gang and non-gang youth (New directions for youth development, 95, 73–99). San Francisco, CA: Jossey-Bass. Swanson, D. P., Spencer, M. B., Harpalani, V., Noll, L., Seaton, G., & Ginzberg, S. (2003). Psychosocial development in racially and ethnically diverse groups: Conceptual and methodological changes in the 21st century. Development and Psychopathology, 15, 743–771. Swanson, D. P., Spencer, M. B., & Petersen, A. (1998). Identity formation in adolescence. In K. Borman & B. Schneider (Eds.), The adolescent years: Social influences and educational challenges: Ninety-seventh Yearbook of the National Society for the Study of Education, Part I (pp. 18–41). Chicago, IL: National Society for the Study of Education. Szalacha, L. A., Erkut, S., García Coll, C., Alarcon, O., Fields, J. P., & Ceder, I. (2003). Discrimination and Puerto Rican children's and adolescents' mental health. Cultural Diversity & Ethnic Minority Psychology, 9(2), 141–155. Telzer, E. H., & Vazquez Garcia, H. A. (2009). Skin color and self-perceptions of immigrant and U.S.–born Latinas: The moderating role of racial socialization and ethnic identity. Hispanic Journal of Behavioral Sciences, 31(3), 357–374. Theokas, C., Almerigi, J. B., Lerner, R. M., Dowling, E. M., Benson, P. L., Scales, P. C., & von Eye, A. (2005). Conceptualizing and modeling individual and ecological asset components of thriving in early adolescence. Journal of Early Adolescence, 25(1), 113–143. doi: 10.1177/0272431604272460 Thompson, M. S., & Keith, V. M. (2001). The blacker the berry: Gender, skin tone, selfesteem, and self-efficacy. Gender and Society, 15, 336–357. Toomey, R. B., & Umana-Taylor, A. J. (2012). The role of ethnic identity on self-esteem for ethnic minority youth: A brief review. Retrieved from http://psycnet.apa.org.proxy.uchicago.edu/index.cfm? fa=search.displayRecord&id=DDC7B74D-CF87–6737–9745– 73D4F09579AF&resultID=25&page=1&dbTab=all&search=true Townsend, T. G., Thomas, A. J., Neilands, T. B., & Jackson, T. R. (2010). I'm no Jezebel; I am young gifted and Black: Identity, sexuality and Black girls. Psychology of Women Quarterly, 34(3), 273–285. doi: 10.1111/j.1471–6402.2010.01574.x Travis, R., & Leech, T. G. (2014). Empowerment-based positive youth development: A new understanding of healthy development for African American youth. Journal of Research on Adolescence, 24(1), 93–116. Tynes, B. M., Umaña-Taylor, A. J., Rose, C. A., Lin, J., & Anderson, C. J. (2012). Online

racial discrimination and the protective function of ethnic identity and self-esteem for African American adolescents. Developmental Psychology, 48(2), 343–355. doi: 10.1037/a0027032 Uhlmann, E., Dasgupta, N., Elgueta, A., Greenwald, A. G., & Swanson, J. (2002). Subgroup prejudice based on skin color among Hispanics in the United States and Latin America. Social Cognition, 20(3), 198–226. Ullrich, H. E. (2010). Is beauty skin deep? The impact of “beautiful attributes” on life opportunities and interpersonal relationships: A tale of two sisters in South India. Journal of the American Academy of Psychoanalysis & Dynamic Psychiatry, 38(2), 243–253. Veale, A. (2006). Child-centered research with ethnic minority populations: Methodological, ethical and practical challenges. Irish Journal of Psychology, 27(1–2), 25–36. doi: 10.1080/03033910.2006.10446225 Veenstra, G. (2011). Mismatched racial identities, colourism, and health in Toronto and Vancouver. Social Science & Medicine, 73(8), 1152–1162. doi: 10.1016/j.socscimed.2011.07.030 Wade, T. J., & Bielitz, S. (2005). The differential effect of skin color on attractiveness, personality evaluations, and perceived life success of African Americans. Journal of Social Psychology, 31, 215 – 236. Wagatsuma, A. (1968). The social perception of skin color in Japan. In J. H. Franklin (Ed.), Color and race (129–165). Boston, MA: Beacon Press. Wallace, S. A., Townsend, T. G., Glasgow, Y. M., & Ojie, M. J. (2011). Gold diggers, video vixens, and Jezebels: Stereotype images and substance use among urban African American girls. Journal of Women's Health, 20(9), 1315–1324. doi: 10.1089/jwh.2010.2223 Way, N., Santos, C., Niwa, E. Y., & Kim-Gervey, C. (2008). To be or not to be: An exploration of ethnic identity development in context. In M. Azmitia, M. Syed, & K. Radmacher (Eds.), The intersections of personal and social identities. New Directions for Child and Adolescent Development, 120, 61–79. Wachtel, T. (2012). Defining restorative. Bethlehem, PA: IIRP Graduate School. Retrieved from http://www.iirp.edu/pdf/Defining-Restorative.pdf Weaver, V. M. (2012). The electoral consequences of skin color: The “hidden” side of race in politics. Political Behavior, 34(1), 159–192. doi: 10.1007/s11109–010–9152–7 Wenzel, M., Okimoto, T. G., Feather, N. T., & Platow, M. (2008). Retributive and restorative justice. Law and Human Behavior, 32, 375–389. Werner, E. E. (1989). High-risk children in young adulthood: A longitudinal study from birth to 32 years. American Journal of Orthopsychiatry, 59(1), 72–81. Wiese, E. B. (2010). Culture and migration: Psychological trauma in children and adolescents.

Traumatology, 16(4), 142–152. Wilder, J., & Cain, C. (2011). Teaching and learning color consciousness in Black families: Exploring family processes and women's experiences with colorism. Journal of Family Issues, 32(5), 577–604. doi: 10.1177/0192513X10390858 Wong, D. S. W., Cheng, C. H. K., Ngan, R. M. H., & Ma, S. K. (2011)). Program effectiveness of a restorative whole-school approach for tackling school bullying in Hong Kong. International Journal of Offender Therapy and Comparative Criminology, 55(6), 846–862. Wood, D., Kurtz-Costes, B., & Copping, K. E. (2011). Gender differences in motivational pathways to college for middle-class African American youths. Developmental Psychology, 47(4), 961–968. Yasui, M., Dorham, C. L., & Dishion, T. J. (2004). Ethnic identity and psychological adjustment: A validity analysis for European American and African adolescents. Journal of Adolescent Research, 19(6), 807–825. Zimmerman, M. A. (2013). Resiliency Theory: A Strengths-Based Approach to Research and Practice for Adolescent Health. Health Education Behavior, 40(4), 381–383. doi: 10.1177/1090198113493782

Chapter 8 Social Inequalities and the Road to Allostatic Load: From Vulnerability to Resilience Robert-Paul Juster, Teresa Seeman, Bruce S. McEwen, Martin Picard, Ian Mahar, Naguib Mechawar, Shireen Sindi, Nathan Grant Smith, Juliana Souza-Talarico, Zoltan Sarnyai, Dave Lanoix, Pierrich Plusquellec, Isabelle Ouellet-Morin, and Sonia J. Lupien Financial support for this work was granted to Robert-Paul Juster by means of a doctoral scholarship from the Research Team on Workplace Mental Health (L'Équipe de Recherche sur le Travail et la Santé Mentale) consortium primarily affiliated with the University of Montreal. Sonia J. Lupien held a Senior Investigator chair on Gender and Mental Health from the Canadian Institutes of Canada Institute of Gender and Health. Thanks are also extended to Jason Berhmann for constructive discussions on social justice. INTRODUCTION Allostatic Load Model SOCIAL INEQUALITIES INFLUENCING ALLOSTATIC LOAD Early Adversity Socioeconomic Status and Health Gradients Race/Ethnicity and Discrimination Shifting Major Economies: Brazilian Context Sex Differences and Gender Diversity Indigenous Peoples of North America and Australia INNOVATIVE BIOCHEMICAL, NEUROLOGICAL, AND COGNTIVE PERSPECTIVES Mitochondria Hippocampal Neurogenesis Cognitive Reserve CONCLUSIONS Clinical Implications Social Policy Implications REFERENCES

Introduction

The developmental trajectories that render individuals vulnerable toward or resilient against diseases of the mind and body are major themes in stress theory, research, and practice. As the field of developmental psychopathology expands our understanding of the multicausal mechanisms involved, vulnerability and resilience are increasingly conceptualized not as static states but rather as probabilistic processes that vary according to the contexts and the consequences investigated (Cicchetti & Toth, 2009; Rutter, 1985; Rutter & Sroufe, 2000). Resilience is therefore defined as a dynamic process that promotes positive adaptation among individuals exposed to severe forms of adversity, stress, and/or trauma (Cicchetti & Garmezy, 1993; Luthar, Cicchetti, & Becker, 2000; Masten, Best & Garmezy, 1990; Rutter, 2012). In order to further identify the intertwined risk factors and protective factors associated with diverse developmental pathways that promote resilience, scientists require multilevel analyses and refined measures of biopsychosocial stress among disadvantaged populations.

Figure 8.1 (A) A transdisciplinary framework in relation to other research frameworks (adapted from Juster et al., 2011). Multidisciplinarity approaches combine knowledge from different disciplines additively. Interdisciplinary approaches involve the appreciation of disciplinary overlap and the integration of discipline-based constructs into higher-order concepts encompassing several disciplines. Transdisciplinarity emerges as a higher-order framework that integrates all knowledge into new concepts and perspectives that encompass complex interactions reaching across disciplinary boundaries. This approach is consistent with perspectives espoused in the field of developmental psychopathology and is best measured by triangulating methods. Adapted from “A transdisciplinary perspective of chronic stress in relation to psychopathology throughout life span development,” Development and Psychopathology, 2011, Volume 23, Issue 3, p. 761, Cambridge University Press. (B) Allostatic load is an exemplar of trandisciplinarism (Juster et al., 2011). Aspects of AL are illustrated here as topic triangles forming a holistic triangle according to diverse disciplines; namely, biological, psychological, sociological, behavioral, and spiritual. As reviewed in Juster, McEwen, and Lupien (2010), each of these life domains has been empirically substantiated in relation to multisystemic biomarkers representing AL that are consequently associated with clinical outcomes. In relation to Figure 8.A, AL is best understood from a transdisciplinary perspective whereby complex interactions among risk factors and protective factors can be identified. See footnote 1. The social determinants of health literature focusing on stress pathophysiology apply increasingly integrative mind-body perspectives to understand how social inequalities influence disease susceptibilities. In this spirit, the neurobiologist Sterling and the epidemiologist Eyer coined the term allostasis to describe dynamic, multifaceted biological processes that maintain physiological stability by recalibrating homeostatic parameters and matching them appropriately to meet environmental demands (Sterling & Eyer, 1988). Analogous to our understanding of resilient systems that have the capacity to dynamically adjust and stabilize systems to adapt to perturbations (Cicchetti, 2013), allostatic processes likewise alter metabolic functioning via compensatory and anticipatory mechanisms for which the brain is the commander in chief. In formulating the concept of allostasis that has since revolutionized our views of what constitutes adaptive and maladaptive stress physiology, Sterling and Eyer (1988) proposed that age and socioeconomic differentials in cardiovascular hyperarousal represented physiological recalibrations to the needs of adversity.

Allostatic Load Model Over the last quarter century, stress researchers have assimilated the concept of allostasis into understanding how biopsychosocial stress gets under our skin and skull (McEwen, 1998a). At a biological level, a stress response exemplifies an allostatic mechanism that integrates mind and body via the brain's perception of environmental stressors (for an earlier review, see Lupien et al., 2006). Under stressful situations, real or interpreted threats to homeostasis initiate a cascade of biological activities that leads to the sympathetic-adrenal-medullary (SAM) axis release of catecholamines such as adrenaline within seconds, followed by the hypothalamic-pituitary-adrenal (HPA) axis production of glucocorticoids such as cortisol

within minutes (Sapolsky, Romero, & Munck, 2000). These activities ultimately serve to adaptively mobilize energy necessary for “fight-or-flight” behaviors (or female-specific “tendand-befriend” behaviors (S. E. Taylor et al., 2000) that will be discussed in a later section on sex and gender diversity) that ensure survival, but at the cost of recalibrating many other biological functions in turn (McEwen & Wingfield, 2003). For instance, compensatory alterations during acute stress include decreased digestion, bodily growth/repair, and reproductive functions counterbalanced to accommodate increased neurological, cardiovascular, respiratory, and immune activities. Although stress responses are adaptive in the short-term, long-term activations can result in the gradual dysregulation of allostatic mechanisms. Shonkoff, Boyce, and McEwen (2009) differentiated stress physiology using a useful taxonomy: (1) positive stress refers to moderate stress responses that are normal, short-lived, and appropriately deactivated; (2) tolerable stress refers to potentially damaging stress responses that get buffered and can even promote protection against future stressors; and (3) toxic stress refers to damaging stress responses occurring under conditions of severe adversity, stress, and/or trauma where there is a lack of internal and/or external resources to cope. In addition, such environmental perturbations can be exacerbated by genetic and epigenetic influences that render brain circuitry inadequate to promote emotional regulation. Under toxic stress, stress hormones become misbalanced and induce an interconnected domino effect on interdependent biological systems that collectively collapse as individual biomarkers topple and trail toward disease. This multisystemic strain attributable to toxic stress is referred to as allostatic load (McEwen & Stellar, 1993). Allostatic load (AL) is defined as the multisystemic wear and tear the brain and body experience when repeated allostatic responses exact their noxious toll under conditions of chronic stress. Over time, chronic stress and AL can perpetuate unhealthy behaviors (e.g., poor sleep, diet, smoking, drinking, physical inactivity) that eventually lead to allostatic overload that in turn culminates in disordered, diseased, and deceased endpoints. In an evolutionary biology framework (McEwen & Wingfield, 2003), allostatic overload comes in two varieties. Type 1 allostatic overload occurs when energy demands exceed energy supply (negative energy balance), for example, when starving or migrating. By contrast, Type 2 allostatic overload occurs when energy supplies exceed energy demand (positive energy balance), for example, when overeating or hibernating. This conceptual distinction is important when we consider the multifaceted interactions between chronic stress and lifestyle choices exercised by people of diverse social strata. Ultimately, allostatic responses necessitate a biological balance that can shift over time in response to adversity that contributes to AL and allostatic overload. Because vulnerability and resilience fall along a continuum (Cicchetti, 2013), applying triangulated methods represents a very powerful means to investigate biopsychosocial stress and AL using transdisciplinary perspectives that transcend disciplinary boundaries (Juster et al., 2011). For example, combining biometrics, psychometrics, and sociometrics together promotes triangulation amenable to both research designs and clinical approaches. The essence of the transdisciplinary tradition is illustrated in Figure 8.1, whereby the antecedents and consequences of AL thus far identified in the literature are holistically represented to

exemplify the importance of comprehensive measurement approaches that collectively assess interconnected domains using multiple disciplinary tools. Beginning with operational definitions and the development of AL algorithms that have proven reliable in identifying developmental trajectories associated with stress-related disease vulnerabilities, the AL model has rapidly evolved and more recently has begun to be applied to detect pathways that promote resilience. Operationalizing Allostatic Load The AL model postulates that stress-related pathophysiology can be detected and objectively measured by assessing a battery of subclinically relevant biomarkers from multiple nonlinear and interconnected biological systems (McEwen & Seeman, 1999). When chronically stressed, the biphasic effects of numerous biomarkers eventually lead to AL and disease in a theoretical sequence embodied in our domino effect analogy: (1) overactivation of primary mediators such as stress hormones and pro- and anti-inflammatory cytokines induce primary effects on cellular activities, leading to secondary outcomes (2), whereby metabolic, cardiovascular, and secondorder immune biomarkers become dysregulated and (3) culminate as tertiary outcomes or clinical endpoints (McEwen & Seeman, 1999). The AL model proposes that by measuring the multisystemic, reciprocal interactions among primary mediators, primary effects, and secondary outcomes, individuals at high risk of tertiary outcomes can be identified (Juster, McEwen, & Lupien, 2010; McEwen, 1998a). This cascade has important implications not only in terms of research but potentially also clinically in detection and prevention of elevated AL; for example, patients with severe mental disorders are prone to developing comorbidities that could be proactively prevented (Bizik et al., 2013). Specifically, physicians routinely incorporate many related biomarkers in typical blood tests; however, attention is largely placed on values reaching clinically significant levels. By including additional biomarkers, identifying preclinical values, establishing clinical norms, and triangulating methods using combined transdisciplinary approaches (e.g., genetic assessment, neuropsychological testing, clinical interviews, psychometrics, sociodemographics), advances in diagnostics, treatment, and prevention can be improved using the AL framework (Juster et al., 2011; Lupien et al., 2006). Pioneering epidemiological work by Seeman and colleagues has led to the development of AL algorithms predictive of numerous tertiary/clinical outcomes and predicted by diverse psychosocial factors. Validation using longitudinal data from the MacArthur Studies on Successful Aging cohort led to a count-based AL index representing the next 10 biomarkers (Seeman, Singer, Rowe, Horwitz, & McEwen, 1997): 12-hour urinary cortisol, adrenaline, and noradrenaline output; serum dehydroepiandrosterone-sulphate (DHEA-S), high-density lipoprotein (HDL), and HDL to total cholesterol ratio; plasma glycosylated hemoglobin; aggregate systolic and diastolic blood pressures (BP); and waist-to-hip-ratio. Participants' values falling within high-risk quartiles (clinical and preclinical ranges based on percentiles) with respect to the sample's biomarker distributions are dichotomized as “1” and those within normal ranges as “0.” Once tabulated, these are summed to yield an AL index ranging from a possible 0 to 10, which can then be used to predict health outcomes. Beyond these traditional

biomarkers, many others have been incorporated into numerous analyses worldwide using similar formulations and/or more sophisticated statistical analyses (e.g., multivariate reduction techniques) that continually refine AL measurement. By using simple or complex AL algorithms, researchers are ever expanding our understanding of the psychosocial antecedents and health consequences of AL. To summarize (see Figure 8.1), increased AL indices are associated with a plethora of antecedents (e.g., socioeconomic disadvantage, poor social networks, workplace stress, maladaptive personality traits, lifestyle behaviors, genetic polymorphisms, etc.) and consequences using cross-sectional and longitudinal study designs (e.g., mortality, cardiovascular disease, psychiatric symptoms, cognitive decline, physical/mobility limitations, neurological atrophy, etc.) (Juster et al., 2010). There are five thematic advantages of applying an elevated-risk-zone system when scoring AL as AL algorithms represent: (1) early-warning signals, since thresholds are anchored at subclinical thresholds; (2) multi-finality, in that similar AL algorithms predict different tertiary outcomes; (3) flexibility, since calculations are based on different biomarker combinations; (4) synergism capturing the cumulative interaction of numerous biomarkers; and finally, most important for this chapter, (5) antecedent adversities predicting individual variation (Singer, Ryff, & Seeman, 2004). In sum, AL algorithms are objective reflections of biological functioning that are intricately interconnected with genetic, neurological, developmental, behavioral, cognitive, and social factors. Modern societies face numerous challenges that are particularly pronounced among those from marginalized strata. In this chapter, we explore how AL disproportionably burdens public health in the developed and developing world. By taking a transdisciplinary perspective of life span development, the first sections of this chapter explore literature focusing on early adversity, socioeconomic status (SES) and health gradients, race/ethnicity and discrimination, shifting economic contexts using Brazil as a case study, the struggles of Indigenous peoples of North America and Australia, and finally the importance of biological sex differences and sociocultural gender diversity in understanding AL. Although the majority of studies focus on risk factors that propagate vulnerability toward AL and pathologies, emerging studies are increasingly identifying protective factors and trajectories that promote resilience, which we believe represents a new frontier.

Social Inequalities Influencing Allostatic Load Many psychosocial stressors that contribute to AL are mirrored throughout the animal kingdom. Allostatic responses are conserved among species and help orchestrate adaptive metabolic changes, for instance, during hibernation or long-distance migration (Gotmann & Wingfield, 2004; Goymann & Wingfield, 2004; Korte, Koolhaas, Wingfield, & McEwen, 2005; McEwen & Wingfield, 2003). Among nonhuman primates and other social species, social hierarchies are also important as dominant and subordinate members experience compromised health depending on the stability of rival competition for higher status (Sapolsky, 2005). Social inequalities therefore exert pernicious effects on health gradients across species.

Viewed from allostatic and AL perspectives, social inequalities are associated with specific physiological phenomena rooted in psychosocial determinants that often have origins in early life experiences. Although dozens of AL studies have linked socioeconomic differences to the prevalence of disease and even death, most reports initially focused on middle-aged and older adults (for reviews, see Beckie, 2012; Dowd, Simanek, & Aiello, 2009; Juster et al., 2011). In the last decade or so, an emerging body of literature has also successfully applied this framework to detect AL levels among disadvantaged children and adolescents exposed to adversity. Many of these have followed in the footsteps of Evans and colleagues (e.g., Evans et al., 1998) by assessing cumulative risk factors, a term that refers to a summary index of early adversities, such as crowding, noise, housing problems, family separation/turmoil, violence, low income-to-needs ratio, single parenthood, and low maternal education.

Early Adversity Individuals exposed to greater social inequalities, such as socioeconomic deprivation, fare worst in terms of physical and mental health (Shonkoff et al., 2009). Nevertheless, our capacity to build on this knowledge to help prevent these negative outcomes is limited. One reason may be that the mechanisms whereby social inequalities underlie long-term vulnerability toward disease remain enigmatic. Specifically, how social inequalities during distinct periods of development shape allostatic systems—especially in early life—are not explicitly encompassed in existing theoretical models. Accordingly, we explore whether allostatic mechanisms are more sensitive to early environmental influences in a manner that affects longlasting changes. Shonkoff and colleagues (2009) proposed that health disparities are rooted in childhood and represent either latent or active sources of vulnerability found at multiple levels of analysis, including physiological systems sensitive to social environments. Akin to the allostasis concept and AL model, these authors postulate that accumulating exposures to adversity—both in terms of quantity or duration—overwhelm the physiological systems designed to support adaptation to stress and induce changes in their physiological set points. This could in turn increase vulnerability to disease in later life stages. In addition, a more profound impact of social inequalities is to be expected, provided they occur during sensitive developmental periods (Shonkoff et al., 2009). The general assumption here is that the earlier the adversity, the greater and more resistant these changes become due to the brain's immaturity and sensitivity to environmental tuning (A. B. Fries, Ziegler, Kurian, Jacoris, & Pollak, 2005). Several lines of research highlight the importance of considering early adversity. This approach may prompt a better understanding of individual differences in later health, which could eventually be further integrated into the AL model. Through the invaluable use of longitudinal studies, epidemiology and developmental sciences have played a major role in substantiating this possibility (Anda et al., 2006; Felitti et al., 1998; Gilbert et al., 2009; Widom, DuMont, & Czaja, 2007). For example, adverse experiences in childhood—but not necessarily in adulthood—predict later psychopathology (Weber et al., 2008). The presumed causal effect of early adversity on later mental health is also supported by rare natural experiments (Costello, Compton, Keeler, & Angold, 2003) as well as experimental designs

controlling for genetic and familial confounds (Arseneault et al., 2008). Thus, robust associations between early adversity and pathology are a starting point to uncovering physiological mechanisms bearing theoretical and potential clinical utility. A second line of research stemming from developmental neurosciences and psychoneuroendocrinology also suggests that early adversity induces persistent differences in primary familial systems. For instance, maternal care in the first weeks of life, but not necessarily afterward, is associated with stable differences in the adult rodents' neuroendocrine and behavioral reactivity to stress (Meaney, 2001). Only indirect evidence suggests that such a period of increased sensitivity to cortisol surges exists in humans, for which the beginning is hypothesized to take place in the first year and persist until puberty (Gunnar & Vazquez, 2006; Neigh, Gillespie, & Nemeroff, 2009). Despite no clear delimitations established among humans (Gunnar & Vazquez, 2006), several brain structures involved in stress regulation are known to remain immature in childhood, adolescence, and early adulthood only to then undergo age-related changes in later life (Lupien, McEwen, Gunnar, & Heim, 2009). As we explore next, the maturity of these structures is particularly vulnerable to the timing and severity of environmental stressors. Life Cycle Model of Stress The brain is the central regulator of allostatic mechanisms that are continuously shaped by the environment. Stress hormones bypass the blood-brain barrier to bind to three key receptordense regions; namely, the prefrontal cortex (PFC), the amygdala, and the hippocampus. Morphological and functional changes in these brain regions occur as adaptations to environmental demands that are essential to understanding pathways toward resilience or vulnerability (Cicchetti, 2013; McEwen, 2009, 2012). Chronic production of glucocorticoids is believed to contribute to frontal lobe malfunctioning, amygdaloidal hypertrophy, and hippocampal atrophy (Lupien et al., 2009). In addition to these neurological processes, individual differences in constitutional (genetics, period of development, environmentally mediated traits and experiences), behavioral (coping strategies and lifestyles), and historical (trauma/abuse, major life events, and stressful environments) factors interactively determine one's sensitivity to chronic stress and AL. Of critical importance is the timing and duration of perturbations, as this profoundly affects neurological development. Lupien and colleagues (2009) have proposed that the consequences of chronic stress and trauma at different life stages depend on which brain regions are developing or declining at the time of exposure. Stress in the prenatal period affects development of the hippocampus, PFC, and amygdala, leading to well-substantiated programming effects. From the prenatal period onward, all developing brain areas are sensitive to the effects of stress hormones; however, some areas undergo rapid growth during key critical windows. For example, from birth to 2 years of age, the developing hippocampus is extremely vulnerable to glucocorticoid surges. By contrast, in late childhood (10 years old), the amygdala appears to be more vulnerable to hypertrophy in the context of lifelong exposure to maternal depression (Lupien, King, Meaney, & McEwen, 2001). During adolescence, the amygdala continues to develop until young adulthood, and finally the frontal lobe undergoes important maturation. Consequently, stress

exposure during this transition into adulthood can have major effects on the PFC. Studies show that adolescents are highly vulnerable to stress because of pubertal changes in gonadal hormones and sensitivities of the HPA axis that can persist into adulthood as potentiation/incubation effects. In middle and older adulthood, the brain regions that undergo the most rapid decline as a result of senescence are once again highly vulnerable to the effects of stress hormones. This period leads to the manifestation of incubated effects from earlier life on the brain referred to as maintenance effects (Lupien et al., 2009). Lifelong brain perturbations can diminish an individual's ability to adapt, leading to subtle recalibrations in stress responsivity that could be used to detect disease trajectories (McEwen, 1998a). According to the life cycle model of stress, regional volumes of these neurological structures in conjunction to biological signatures (e.g., hypercortisolism or hypocortisolism) can be used to predict differential risk profiles for specific psychopathologies (e.g., depressive or posttraumatic stress disorder) in adulthood as well as inform when certain traumas might have occurred in early life (Lupien et al., 2009). This hypothesis complements the AL model. Given that direct measurement of central nervous system substrates are costly and potentially invasive, indirect assessment using peripheral biomarkers routinely collected in blood draws could be used to measure AL algorithms (Juster & Lupien, 2012a). By this same logic, developmental trajectories that promote resilience at a neurological level could also be identified. Neuroendocrine responses triggered by early adversity are thought to play a pivotal role in the life cycle model of stress and the AL model. The concentration of glucocorticoid receptors (GR) in the PFC, amygdala, and hippocampus (Sapolsky, 2000) are consistent with the proposed mediating pathways translating exposure to early adversity into lasting changes in perceiving, responding, and coping with stressors (e.g., health behaviors such as sleep, diet, and exercise) as well as nonrandom exposures (e.g., distinct living environments, such as crowded, unattractive, dangerous, noisy, and polluted spaces). The next three subsections provide brief overviews of how changes in cortisol, immune, and epigenetic mechanisms following early adversity may become biologically embedded (Hertzman, 1999) to render individuals vulnerable to future pathologies. Consistent with the life cycle model of stress, such processes can affect the long-term transaction between individuals' (mal)adaptation to their environment. Hypothalamic-Pituitary-Adrenal Axis Evidence from animal and human studies demonstrate that the HPA axis is integral to the physiological and behavioral adaptation required to face stressors (Sanchez, 2006; Tarullo & Gunnar, 2006). However, significant associations have been documented between a wide range of adverse life circumstances and distinct cortisol profiles (e.g., cortisol awakening response; Engert, Efanov, Dedovic, Dagher, & Pruessner, 2011) and disrupted diurnal rhythms (Bruce, Fisher, Pears, & Levine, 2009; Cicchetti & Rogosch, 2001; Cicchetti, Rogosch, Gunnar, & Toth, 2010; Lupien et al., 2001; Neigh et al., 2009); for review, including conflicting findings, see Dowd et al. (2009) and Tarullo and Gunnar (2006). Early adversity is also linked to disrupted patterns of cortisol reactivity to stress, manifested in higher (Heim et

al., 2000) and lower responsivity (Carpenter et al., 2007; Lovallo, Farag, Sorocco, Cohoon, & Vincent, 2012). More recently, early life stress has also been linked to higher cortisol levels in hair. Hair cortisol is an increasingly validated matrix to monitor cumulative cortisol secretion retrospectively (Kirschbaum, Tietze, Skoluda & Dettenborn, 2009). For example, reports of higher hair cortisol have been shown among hospitalized newborns (Yamada et al., 2007), preschoolers growing up with less educated parents (Vaghri et al., 2013), as well as young adults reporting major stressors (e.g., death of a relative) or psychological problems (Karlen, Ludvigsson, Frostell, Theodorsson, & Faresjo, 2011). Similar to findings in regard to salivary cortisol (J. H. Fries, Hellhammer, & Hellhammer, 2005), prolonged stress has also been associated with lower hair cortisol in young adults (Stalder et al., 2012), not just with higher levels. Concurrent mental health as well as the nature, timing, and prolonged adversity have been hypothesized to explain these seemingly opposite patterns of cortisol secretory patterns (Gunnar & Vazquez, 2001; Miller, Chen, & Zhou, 2007). Direct tests of these hypotheses are unfortunately scarce in humans (Trickett, Noll, Susman, Shenk, & Putnam, 2010) and should be the focus of continued transdisciplinary research. Timing of stressor exposure is essential to understanding the heterogeneity of future HPA axis functioning. Preliminary evidence suggests that adversity taking place before age 11 (including pre-/postnatal) is associated with higher cortisol responsivity during psychosocial stress whereas adversity experienced after age 11 is linked to lower levels (Bosch et al., 2012). Future studies with repeated measures of adversity collected prospectively, beginning at birth and persisting into adolescence, could further our understanding of these time-sensitive trajectories hypothesized to differentially affect HPA axis functioning. Such studies could include, but are not restricted to: the timing when adversity occurs; stress-sensitization effects of pre-/postnatal adversity when adversity is again experienced at later time points (Bosch et al., 2012); the length of time since the adversity ended (Miller et al., 2007; Trickett et al., 2010); differential maturational processes of brain structures regulating reactivity to stress (Lupien et al., 2009); and age- and puberty-related changes in HPA axis functioning (Dahl & Gunnar, 2009; Shirtcliff et al., 2011; Stroud et al., 2009; Trepanier et al., 2013). Several considerations should be noted to guide future directions. Studies that link early adversity to HPA axis functioning are often constrained by a limited number of cortisol samples or other biomarkers during restricted periods of development. Moreover, study designs are generally correlational in nature and only partially address confounders. Intervention studies (often quasi-experimental) have shown that improvement (or alleviation) of detrimental proximal environmental influences help to normalize HPA axis functioning (Fisher, Gunnar, Dozier, Bruce, & Pears, 2006). Yet it remains unclear whether the same mechanisms are at play in the context of early adversity and whether these pervasively exert a causal effect on AL primary mediators. For ethical reasons, studies involving manipulations of these variables are not possible to conduct among humans. Nevertheless, complementary designs could be used (Rutter, 2007), including the investigation of teratogenic prenatal influences in the context of in vitro fertilization and discordant monozygotic (MZ) twin designs.

In the context of peer victimization, distinct patterns of cortisol reactivity to stress in a subsample of identical twins discordant for bullying victimization from the Environmental Risk Longitudinal Twin Study is consistent with a causal impact of early adversity on cortisol reactivity to stress (Ouellet-Morin et al., 2011). Specifically, nonbullied twins showed increases in cortisol following social stress whereas bullied cotwins did not exhibit this increase. Because of the study design, the difference in neuroendocrine response to stress cannot be explained by the genetic background, shared familial influences, or other individual experiences or characteristics. Since these factors cannot be disassociated, this suggests that bullying victimization exerts an environmentally mediated effect on the HPA axis. More generally, inadequate HPA axis responsivity in the context of adversity may indicate that these experiences have become biologically embedded and thus more resistant to positive changes of interventions. Finally, few studies have examined prospectively whether neuroendocrine changes triggered by early adversity persist beyond adolescence. Preliminary evidence of a flattened pattern of morning cortisol secretion found in middle-age adults with a history of childhood maltreatment ascertained prospectively suggests that early adversity indeed biologically embeds future biological functioning (Power, Thomas, Li, & Hertzman, 2012). Future studies including repeated measures of adversity and HPA axis functioning collected prospectively across the life span will contribute to better understanding of the roots of health disparities. Immune Functioning In addition to investigations of HPA axis functioning, research indicates that inflammation is associated with an increased risk of pathology, including cardiovascular diseases, autoimmune diseases, obesity, and affective disorders (Buckley, Fu, Freeman, Rogers, & Helfand, 2009; Howren, Lamkin, & Suls, 2009). Although acute stress enhances adaptive immune functions via cortisol and catecholamines, chronic stress and poor health behaviors ultimately suppress adaptive immunity and further lead to a chronic inflammatory state as immune mediators cease to promote innate immune defenses and adaptive immunity (Dhabhar, Malarkey, Neri, & McEwen, 2012). Interest in the early origins of individual differences in inflammation and its hypothesized consequences on health has recently grown and could constitute another pathway linking early adversity to later pathology (Felitti et al., 1998; Green et al., 2010). In a series of studies, Danese and his team (2009) reported persistent changes in the innate immune system of children exposed to early adversity. They reported higher levels of inflammation in 32-year-old adults who were maltreated or socially isolated as children in comparison to controls. Such experiences accounted for approximately 13% of the cohort with elevated inflammation (high sensitivity c-reactive protein [CRP] > 3 mg/L). Other longitudinal studies corroborate the hypothesized impact of early adversity persisting into adulthood, as associated with CRP (S. E. Taylor, Lehman, Kiefe, & Seeman, 2006), interleukin-6 (IL-6), and fibrinogen (Slopen et al., 2010). Likewise, among older adults these associations persist, as suggested by higher levels of proinflammatory cytokines IL-6 and tumor necrosis factor-α (TNF-α) in caregivers and noncaregivers with a history of childhood abuse (Kiecolt-Glaser et al., 2011). Interestingly, these patterns of inflammation were also detected in the context of

acute stress in both a standardized protocol and a natural setting (Carpenter et al., 2010; Gouin, Glaser, Malarkey, Beversdorf, & Kiecolt-Glaser, 2012). Additional studies also suggest that the impact of early adversity on inflammation, including socioeconomic factors, may already be detectable in late childhood (Danese et al., 2011) and adolescence (Miller & Chen, 2010). Higher CRP levels were reported in children living in Indigenous Canadian communities; this finding represents a higher risk of impending cardiovascular problems and further supports the idea that the early roots of adult diseases implicate inflammatory mediators (for an extended review, see (Danese & McEwen, 2012; Fagundes, Glaser, & Kiecolt-Glaser, 2013). Beyond primary mediators representing neuroendocrine and immune parameters, increasing attention has focused on genetic propensities in interaction with environmental triggers. In the same spirit as recent studies of the HPA axis, future studies should explore whether individual (genetic background) and stress-related factors that affect stress responsivity, AL, and disease trajectories. Deoxyribonucleic Acid Methylation In the last decade, innovative experiments using animal models of early-life stress have shown that the genome is not as immutable as previously thought. Rather, it responds dynamically to the environment through the regulating influence of epigenetic processes (Bagot & Meaney, 2010). Epigenetics refers to a series of modifications in the chemistry of deoxyribonucleic acid (DNA) and its associated proteins (collectively termed chromatin) that have the potential to alter expression without changing its sequence. The epigenome is conceptualized as an interface between the environment and the genome. In addition to the intrauterine environment, recent evidence suggests that postnatal social environments can also alter epigenetic signals (Szyf, McGowan, & Meaney, 2008), which represent another biological mechanism by which childhood adversity could have long-lasting influences on health (Jirtle & Skinner, 2007; Shonkoff et al., 2009; Tsankova, Renthal, Kumar, & Nestler, 2007). To date, most empirical findings supporting the epigenome construct come from animal models of early-life stress. Specifically, maternal care received by rat pups during the first week of life changes DNA methylation of the GR gene promoter in the hippocampus, GR expression, and stress reactivity in adulthood (Meaney & Szyf, 2005). DNA methylation is one of several epigenetic mechanisms but is often regarded as the most stable and representative (Talens et al., 2010). Additional studies have since extended these findings to other genes regulating stress and neurodevelopment, such as the arginine vasopressin (AVP) and brain-derived neurotrophic factor (BDNF) genes (Murgatroyd et al., 2009; Roth, Lubin, Funk, & Sweatt, 2009). The impact of early stress on DNA methylation has also been documented in nonhuman primates (Kinnally et al., 2010; Provencal et al., 2012). Comparatively less research has investigated the epigenetic remodeling following early adversity in humans. Preliminary reports of higher methylation of the GR gene and the serotonin transporter (SERT) gene in adults with a history of childhood abuse or low SES (ascertained retrospectively) are consistent with animal data supporting the epigenome (Beach, Brody, Todorov, Gunter, & Philibert, 2010; Labonte et al., 2012). Similarly, differential

clustering patterns of DNA methylation have been reported across the genome in children raised in institutional care (Naumova et al., 2011). Birth cohorts with prospective measures of adversity have recently contributed to clarifying the temporal sequence of these associations and reported distinct DNA methylation profiles in adults exposed to different levels of socioeconomic adversity during childhood (Borghol et al., 2012). To the best of our knowledge, only two longitudinal studies have explored the associations between early adversity and DNA methylation patterns with diverse measures of adversity (not only SES) collected prospectively throughout childhood (Essex et al., 2011; Ouellet-Morin et al., 2013). The first study comprises 109 adolescents exposed to maternal stress during their first year who showed elevated DNA methylation in 139 sites located across the genome (Essex et al., 2011). The second study showed that bullied twins had higher SERT DNA methylation compared to their nonbullied MZ cotwins, again supporting an environmentally mediated effect of childhood adversity (Ouellet-Morin et al., 2013). Childhood adversity is theorized to differently affect DNA methylation depending on the developmental periods when these experiences occur (Roth & Sweatt, 2011). Preliminary findings in humans tend to support this hypothesis. For instance, early adversity affects DNA methylation around the time of conception but not in late gestation for famine exposure; in the first year of life but not during the preschool years for maternal stress; and in the second but not the third trimester for maternal depressed mood (Devlin, Brain, Austin, & Oberlander, 2010). By and large, the understanding of the role of DNA methylation is constrained by the few studies conducted in humans and the lack of repeated measures of adversity and DNA methylation across development. This makes it difficult to establish clear-cut temporality and thus to evaluate whether the time when adversity is experienced should be a concerted focus in integrated revisions of the AL model. Moreover, the role of DNA methylation within the GR promoter (or other genes directly or indirectly affecting stress regulation) and the underlying biological pathways by which these changes affect GR expression remains poorly understood in humans. The studies reviewed are nevertheless crucial to our continued understanding of how DNA methylation could eventually be integrated in the allostasis and AL framework, as GR plays a key role in adaptive functions as well as pathophysiology (namely by turning on and turning off the HPA axis). Future research in this promising area also will contribute to existing frameworks delineating the mechanisms by which early adversity exerts long-lasting influences on physical and mental health. Akin to emerging epigenetic studies in humans, AL studies have also revealed intergenerational associations of lower SES. For example, non-Hispanic Caucasian and African American children with less educated parents have higher AL levels, driven heavily by insulin dysregulation (Goodman, McEwen, Huang, Dolan, & Adler, 2005). Within an allostatic framework, insulin dysregulation is part of a broader network of disrupted allostasis related to cortisol, inflammation, and a myriad of processes that differ markedly as a function of social inequalities. In the next section, we explore literature linking AL indices to SES mostly in adulthood but also increasingly at earlier life stages.

Socioeconomic Status and Health Gradients

SES represents education, income, neighborhood, and occupational features that interact strongly with AL throughout life (Szanton, Gill, & Allen, 2005). The MacArthur Study of Successful Aging used to validate the AL model consistently found that higher AL predicts incidences of cardiovascular disease, physical declines, cognitive impairments, and all-cause mortality (Seeman et al., 1997; Seeman et al., 2004a; Seeman, McEwen, Rowe, & Singer, 2001b). Interestingly, even in this representative group of successful agers from diverse race/ethnicity and SES, lower education was a powerful contributor to higher AL levels (Seeman et al., 2004a). This patterning provides evidence of health gradients found not only between but also within socioeconomic strata (Adler & Snibbe, 2003). Likewise, a cross-cultural study of Taiwanese elders from the Social Environment and Biomarkers of Aging Study (SEBAS) revealed that higher AL as well as lower education and income were independently associated with health status (P. Hu, Wagle, Goldman, Weinstein, & Seeman, 2007). By combining the Taiwanese and MacArthur cohorts, lower SES (lower education, occupational status, and finances) together with greater social challenges (e.g., recent widowhood, high family demands) related to higher AL (Weinstein, Goldman, Hedley, Yu-Hsuan, & Seeman, 2003). This finding suggests that SES and adversity might synergize among different older adults from North America and Asia. Some of the most compelling epidemiological evidence in support of the AL index has come from the large-scale American National Health and Nutrition Examination Survey (NHANES). Findings show that AL mitigates the effects of education and income gradients predictive of ischemic heart disease and periodonal disease (Sabbah, Watt, Sheiham & Tsakos, 2008). Tracking about 15,000 participants from the same cohort over 6 years, it was shown that those living in poverty show sharper increases in AL up until middle age and then plateau around ages 70 and above (Crimmins, Johnston, Hayward, & Seeman, 2003). Strikingly, those with high AL have a life expectancy that is 6 years shorter than those with low AL but with similar poverty status and matched for sex (Crimmins, Kim, & Seeman, 2009). Beyond individuals and households, AL studies on SES have also begun assessing neighborhood characteristics. A cross-sectional analysis of 12- to 20-year-olds from NHANES assessing multiple levels nested at the levels of individuals, families/households, and census tracts elegantly demonstrated gradients in AL levels that increased incrementally among adolescents living under low-, medium-, to high-cumulative-risk neighborhoods (Theall, Drury & Shirtcliff, 2012). Likewise, a large-scale multiethnic analysis revealed that Detroit-area neighborhood poverty was positively associated with AL independent of household income and covariates and was furthermore mediated by perceived environmental stress within one's neighborhood (Schulz et al., 2012). Socioeconomic factors may affect AL differently according to one's sex and age. A recent study of 199 healthy Montreal workers revealed that higher occupational status (classified using a pure-type scaling system based on median education and income in a given occupation) was associated with higher AL in men, whereas the reverse occurred for women (Juster, Moskowitz, Lavoie, & D'Antono, 2013a). As has been substantiated in primate research that parallels socioeconomic inequalities in humans (Sapolsky, 2005), males at the top of the

hierarchy might be challenged by unique sets of adversity that exacerbate rather than attenuate physiological functioning. Specifically, socially dominant male primates experience the highest physiological dysregulations when reigning in unstable hierarchies in which subordinates continuously challenge their authority (Sapolsky, 2005). This finding has been confirmed in a comparative study of social status, AL, and glucocorticoid profiles among free-ranging animals (Goymann & Wingfield, 2004). When translated to humans, it is plausible that this pattern also exists among men occupying higher occupational status at the price of higher AL. Given the worldwide changes in wealth and health as a function of women further occupying the workplace, it will be important for future studies to also understand the experiences and health of women in higher occupational statuses who have so often faced unique challenges, unknown to men, to reach the top. Resilient “Shift-and-Persisting” and Life Histories AL studies identifying pathways toward resilience among low-SES individuals are emerging. Among women of different socioeconomic strata, workplace stress moderates AL levels: Those with lower occupational status, but who have higher decision latitude or job control at work, experience lower AL levels (Juster, Moskowitz, Lavoie, & D'Antono, 2013a). In the Biomarker Sub-Study of the Study of Midlife in the United States (MIDUS) cohort, it has been proposed that individuals of either sex who are able to “shift-and-persist” out from low childhood SES might be more resilient to AL. This survival strategy is of particular importance to individuals faced with social adversity, as it combines shifting one's cognitive appraisals and emotional regulation positively to adapt to stressors while persisting by maintaining hopes for the future (Chen & Miller, 2012; Chen, Miller, Lachman, Gruenewald & Seeman, 2012). In addition, increased job control may be another way in which positive working conditions can moderate the otherwise negative effects of lower SES. Within the framework of the shift-and-persist hypothesis, we propose that protective workplace characteristics represent one avenue whereby individuals can promote their health even when unable to shift into higher SES boulevards. Life histories in socioeconomic mobility and social and family relationships are increasingly explored in the AL literature. Using the Wisconsin Longitudinal Study, Singer and Ryff (1999) have applied such pathways and demonstrated that childhood parental bonding and household income categories are associated with AL. Contrasting childhood context with adult finances and relations revealed that increased downward mobility—that is, a depreciation in one's SES over time—was related to increased AL, as was low parental income coupled with negative social relationships. Most important, this approach can be used to identify resilient pathways whereby, for example, economically disadvantaged individuals can compensate by developing positive social relationships that buffer against AL (Singer & Ryff, 1999). This approach can also be used to track the pervasive effects of adversity that accumulate throughout life. Concordantly using the MIDUS cohort, individuals who grew up in low-SES contexts (low parental education, income, and welfare status) and who remained in disadvantaged circumstances in middle and later adulthood (indexed according to educational attainment, household income, difficulty paying bills, availability of money to meet basic

needs, and current financial situation) showed the highest AL levels, followed by individuals classified as downwardly mobile, upwardly mobile, and individuals who always had high SES (Gruenewald et al., 2012). Although these results may be related to elusive critical thresholds in developmental processes, the authors stated that their findings supported the cumulative risk hypothesis that accumulated SES adversity accelerates biological damage. As we explore next, cumulative risk and differences in SES are more pronounced among marginalized groups within society.

Race/Ethnicity and Discrimination Gradients in AL related to SES are compounded by race/ethnic disparities. For instance, African Americans with peripheral arterial disease in the NHANES had the highest AL levels compared to other race/ethnicities (Nelson, Reiber, Kohler, & Boyko, 2007). In another multiethnic analysis of NHANES, African Americans of both sexes had higher AL levels than matched Caucasians, partially explaining a greater overall risk for cardiovascular and diabetes-related mortality after adjusting for education, poverty status, and health insurance status (Duru, Harawa, Kermah, & Norris, 2012). It has been proposed that African American women are particularly vulnerable as they show the most consistently elevated levels of AL across age groups. Moreover, neighborhood characteristics and race/ethnic segregation were found to be particularly damaging to African American women in terms of their increased rates of obesity (Chang, Hillier & Mehta, 2009). The “weathering hypothesis” states that the health of non-Caucasian women deteriorates in early adulthood as the physical consequences of cumulative socioeconomic deprivation manifest themselves (Geronimus, Hicken, Keene, & Bound, 2006). This chronic strain in turn contributes to intergenerational effects that complicate reproductive activities and lead to decrements in health at greater rates than among other more privileged societal groups. In this vein, another NHANES analysis of women between the ages of 17 and 30 revealed that those with high AL levels were two times more likely to have experienced early menarche (10 years old or earlier); early menarche was in turn more common among women who were African American, less educated, currently poorer, smoked, and/or had a history of depression (Allsworth, Weitzen, & Boardman, 2005). Consistent with this sex-specific vulnerability, data from the Jackson Heart Study found that AL indices and constituent biomarker clusters are more consistently dysregulated among African American women, but less so among African American men (Hickson et al., 2012). Adding insult to injury, African American women who have had breast cancer also experience higher AL levels; however, in the NHANES cohort, Caucasian women do not (Parente, Hale & Palermo, 2013). Beyond the weathering hypothesis, the deleterious interaction effects of race/ethnicity and SES factors can be amplified by psychosocial adversities, such as discrimination or acculturation (Chapman et al., 2009; Friedman, Williams, Singer, & Ryff, 2009; Kendall & Hatton, 2002; Mays, Cochran, & Barnes, 2007; Myers, 2009). For instance, social isolation, single motherhood, early parenthood, and numerous other factors must be further identified as effect modulators of sex and race/ethnicity specific associations. Moreover, although the influence of discrimination is difficult to quantify, a report suggests that in adolescence, the association

between poverty and elevated AL is mediated by perceived discrimination (Fuller-Rowell, Evans, & Ong, 2012). Notably, impoverished Caucasians are less likely than nonimpoverished African Americans to have higher AL. For individuals of lower neighborhood SES in the NHANES cohort, the risk of having the highest AL levels is 200% for African Americans, 70% for urban Mexican Americans, and 30% for Caucasians (Merkin et al., 2009). Beyond objective sociometrics, it is possible that many of these findings are driven by interaction effects with subjective factors like discrimination that are experienced in distinct ways. For instance, U.S.-born Mexican Americans and African Americans show higher levels of AL than foreign-born Hispanics and Caucasians, suggesting that migration interacts with SES effects in compromising immigrants' health and well-being (Seeman et al., 2008). Regardless of racial/ethnic stratifications, however, lower education and income are consistently associated to AL. A recent cross-sectional analysis of NHANES confirmed that low-SES adolescents and African American adolescents experience elevated AL levels; however, the relative advantage of Caucasians and Mexican Americans in comparison to African American adolescents diminishes around age 15, at which point metabolic markers increase specifically among Mexican American (Rainisch & Upchurch, 2013). Resilient “John Henryism” and Familal Support In terms of potential resilient pathways, a fascinating research program developed by Brody and colleagues (2013b) has investigated the John Henryism theory among African American youth. Legend says that John Henry was an industrious African American railroad worker in the late 1800s who challenged and defeated a steam-powered mechanical drill in a steeldriving contest, but died thereafter of exhaustion. The John Henryism theory (James, 1994) therefore refers to behaviors that involve high-effort coping and hyperarousal that fuel singleminded mental concentration and drive physical energy needed to succeed when faced with overwhelming odds (Brody et al., 2013b). Resilience may thus be seen in key life domains (e.g., overt behaviors, academic performance), but it comes at a physiological price. This proposition was tested among a cohort of rural African American youth from the southern United States who had cumulative risk factors (higher scores indicating lower SES) and teacher-reported issues wih self-control and competence measured from ages 11 to 13. At follow-up when 19 years old, depressive symptoms, externalizing behaviors, and AL levels (overnight urinary cortisol, adrenaline, and noradrenaline; resting diastolic and systolic BP; and body mass index [BMI]) were measured. Results revealed that preadolescents from lowSES environments with high competence displayed low levels of teacher-reported adjustment problems, but at the cost of higher AL at age 19. These findings support the John Henryism theory and sadly suggest that resilience may be skin deep (Brody et al., 2013b). Said otherwise, adaptation in domain-specific areas may come at a physiological cost. This finding highlights the notion that resilience is not a status or a generalized state of global “resiliency” but rather a constellation of adaptations to life's challenges in terms of both positive and negative domains. In a separate investigation of genetic and contextual associations using the same cohort, a

three-way interaction effect was found for environmental adversity (low family support) with genetic susceptibilities (e.g., short allele of serotonin transporter-linked polymorphic region genotype) predicted AL levels in early adulthood (Brody et al., 2013c). In the hopes of identifying different developmental trajectories, another analysis also using genetic information as well as psychological adjustment between ages 14 and 18 applied latent profile analysis to demarcate two developmental profiles: (1) a physical health vulnerability profile characterized by high cumulative SES risk/high AL/low adjustment problems; and (2) a resilient profile characterized by high cumulative SES risk/low AL/low adjustment problems (Brody et al., 2013a). While carrying the s allelle for the 5-HTTLPR genotype (diathesis) and receiving less peer support (stress) predicted membership in either profile, so did protective parenting and emotional regulation. Brody and colleagues argued that while resilience may be skin deep, the positive influence of family support can help “thicken” skin regardless of color and render protection against genetic susceptibilities and adverse SES conditions. More broadly, parental support appears to be a powerful pathway toward long-standing protection against elevated AL. An analysis of the 27-year longitudinal northern Swedish cohort showed that high parental involvement in children's academic studies (measured by teacher and pupil ratings) at age 16 was associated with low AL (16 biomarkers) at age 43 while accounting for an array of life course sociodemographics (Westerlund, Gustafsson, Theorell, Janlert, & Hammarstrom, 2013). Taken together, this impressive body of work is consistent with findings on resilience in the field of developmental psychopathology more broadly. This fact attests to the vast heterogeneity and multifinality manifested by differential configurations of genetic processes and contextual challenges (Cicchetti et al., 2010). Moreover, these findings also speak to the importance of race/ethnic, life experiences, and genetic interactions in explaining pathways toward vulnerability or resilience that can be modulated by modifiable behaviors as simple as parental involvement in scholastics.

Shifting Major Economies: Brazilian Context Contextual factors beyond SES and race/ethnicity influence health outcomes internationally. It is important to highlight that the majority of studies linking SES to stress pathophysiology (Cohen et al., 2006; Hajat et al., 2010; Lupien et al., 2001) are based on samples from the developed world. Despite such adversity, low-SES individuals from the developed world are generally living in better conditions than those of comparative SES living in the developing world. This fact presents an important interpretive issue when we consider that 80% of Earth's population lives in developing countries, where the majority of chronic disease burden for conditions like hypertension and depression occur (Ibrahim & Damasceno, 2012; World Health Organization [WHO], 2012). In this section, we substantiate the importance of context in relation to stress-related diseases using Brazil's shifting socioeconomic conditions as a case study and as a cross-cultural contrast to the Canadian context. Social inequalities related to unemployment, deprivation, and income disparities are pronounced in countries with emerging economies that further widen gaps between social classes (WHO, 2012). As elsewhere, those at the bottom of the social hierarchy consistently demonstrate poorer health than those at the top (Demakakos, Nazroo, Breeze, & Marmot, 2008;

Kawachi & Kennedy, 2002). In particular, subjective feelings related to unfavorable social comparisons—in other words, feeling poor (subjective SES)—may further explain why low SES predicts poorer health (Adler, Epel, Castellazzo & Ickovics, 2000; Sapolsky, 2005; Wilkinson, 1996). Ultimately, the large gap between social positions strongly promotes discrimination (Jackman, 1994), which intensifies the consequences of poverty. Perceived discrimination—racial or otherwise—is associated with stressful life experiences and adverse health outcomes (Clark, Anderson, Clark, & Williams, 1999; Dion, 2001; Harrell, Hall, & Taliaferro, 2003; Harrell, Merrit, & Kalu, 1998). For example, perceived unfairness, racial discrimination, and nonracial discrimination are positively associated with coronary calcification (Lewis et al., 2006), and are independent predictors of coronary events (De Vogli, Ferrie, Chandola, Kivimaki, & Marmot, 2007). Although these findings are based on minority groups living in developed countries, discrimination is an especially important contextual factor in countries with emerging economies. Indeed, discrimination generates social inequalities that transcend group membership(s) as part of minority or majority groups. For example, native Africans living in South Africa exhibit higher levels of psychological distress than the Caucasian minority, with perceived chronic discrimination positively associated with distress (Williams et al., 2008). In contrast, social support and social integration are protective factors that are each mitigated by different mechanisms (Cohen, 2004). Although social support attenuates stress through psychological and material resources provided by social relationships, social integration influences health based on the participation of the individual in the social relationships and the sense of belonging this creates (Brissette, Cohen, & Seeman, 2000; Cohen, 2004). Social supports are coping strategies that dampen stress by (a) altering the appraisal or perception of a potential threat and (b) modulating reappraisals of the potential threat after exposure (Cohen & Wills, 1985). First, stress reactivity might not be activated as intensely among individuals who feel supported. Indeed, social support seems to benefit health only for individuals under adverse circumstances since there is no effect in the absence of highly stressful demands (Cohen, 2004; Cohen, Gottilieb & Underwood, 2000). Second, after perceiving an event as threatening, individuals with greater social support show enhanced emotional regulation and reduced physiological responsivity. Several studies demonstrate that high social support corresponds to lower cortisol levels in comparison to those with low social support (Eisenberger, Taylor, Gable, Hilmert, & Lieberman, 2007; Rosal, King, Ma, & Reed, 2004; Uchino, Cacioppo, & Kiecolt-Glaser, 1996). In the context of countries with emerging economies, we know of only one study showing that individuals with higher social support display lower BP as well as self-report less perceived stress and depressive symptoms compared to those with limited or no access to social support networks (Dressler, Balieiro, & Dos Santos, 1997). Regarding social integration, one's active engagement within social networks can promote positive psychological health and can facilitate health-promoting behaviors. In contrast to social support, the benefits of social integration on health are independent of whether an individual experiences adversity or not (Cohen, 2004). Interestingly, greater social integration is associated with lower rates of mortality (Berkman & Glass, 2000) as well as lowered risk of cancer recurrence,

psychopathology, and cognitive decline with aging (Cohen et al., 2000). In sum, these findings suggest that psychosocial factors, such as negative feelings related to unfavorable conditions, perceived discrimination, social support, and social integration, represent significant risk factors or protective factors associated with stress and health. Within the frame of advancing the AL model at an international level, these findings in turn raise questions: for example, do contextual differences explain discrepancies in perceived stress and physiological functioning between nations with developed and developing economies? Psychosocial Stress: Canadians versus Brazilians A recent cross-national collaborative study conducted between our Canadian and Brazilian research groups revealed significant differences in self-reported psychosocial stressors as well as differential patterns of basal and stress-reactive cortisol profiles. A rich literature on the psychosocial determinants of neuroendocrine responses reveals that situations that are novel, unpredictable, threatening to the self, and uncontrollable additively contribute to stress hormone responses (Dickerson & Kemeny, 2004; Lupien et al., 2006; Mason, 1968). With this in mind, 1,156 healthy youngsters, adults, and elderly participants from Canada and Brazil— developed and emerging economies, respectively—were asked to indicate among a list of nine different factors (novelty, unpredictability, low sense of control, social evaluation, time pressure, work overload, conflict, unbalance and children) which factors elicited the greatest distress. For Canadians, “children,” “novelty,” and “unpredictability” were the major stressors; Brazilians indicated “time pressure,” “work overload,” and “conflict” as their major stressors. Differences regarding education level, SES, work organization, and experiences may synergistically influence these perceptions and in part explain the discrepancies between Canadian and Brazilian contexts. In particular, Brazilians experience long workdays (between approximately 40 to 44 hours a week on average), work overload (43.6% of Brazilians work overtime), informal sector or under-the-table employment (10.7% of Brazilians work without the coverage of labor laws and employment insurance), and low payments associated with low purchasing power (Institute for Applied Economic Research, 2009). In addition, Brazilian inhabitants must routinely face such challenges as unequal access to public health services, education, and transportation as well as frequent exposure to violence, overcrowding, and traffic. These contextual factors explain the differences in stressor sensitivities between Canadians and Brazilians. Beyond the geographical and contextual factors that influence perceived distress, people from different environments may present different patterns of physiological stress responsivity. To the best of our knowledge, only a handful of studies have investigated cortisol profiles among nondeveloped nations living in disadvantaged conditions (Decker, 2000; Flinn & England, 1997; Nyberg, 2012; Pike & Williams, 2006) and demonstrated both hyper- and hypocortisolemic profiles; therefore, any conclusions at this time are limited. None of these studies, however, compared stress response profiles between countries, each with its own distinct cultural, economic, political, and social realities. Continued cross-cultural

investigations will help the scientific community and health organizations better understand pathways toward vulnerability or resilience in diverse and ever-changing contexts. In delineating cross-cultural patterns linking stress to disease, factors related to migration, immigration, and nativity are important to consider. To date, two cross-cultural studies have shown that difficulty verbalizing (Kim, 2008) and race/ethnicity (Hajat et al., 2010) represent factors associated with both low and high cortisol levels among adult immigrants in contrast to native inhabitants. This finding may, however, be due to acculturation biases that may underestimate the differences observed. For example, higher basal cortisol levels were also observed in Australian Aboriginal communities compared to British inhabitants from Oxford (Harrison, 2001). Conversely, mean basal cortisol levels of Tsimane Amazonian adult foragers were lower compared to American participants from the Coronary Artery Risk Development in Young Adults (CARDIA) study (Cohen et al., 2006; Nyberg, 2012). In addition to these findings among different international communities, Kim (2008) demonstrated an influence of culture on cortisol response to a speech task comparing European Americans with Asian Americans. In sum, nationality and inherent sociocultural factors related to race/ethnicity and acculturation influence stress. Although Canadians and Brazilians share Western cultural roots, they differ vastly with respect to urban conditions, physical conditions (e.g., climate/light exposure, noise, pollution, crowding), history, sociocultural views, and organizational values that may ultimately modulate their respective sensitivities and HPA axis adaptation to daily stressors with important health disparities. For instance, a WHO report showed 39.4% prevalence of raised BP among male adults aged ≥ 25 years among Brazilian inhabitants (WHO, 2012). Similarly, higher rates of diabetes (421 cases per 100,000 population by case) and unipolar depressive disorders (1,410 cases per 100,000 population by case) were reported among Brazilian inhabitants. These findings collectively suggest that chronic stress exposure, cortisol dysregulation, and AL are heightened among inhabitants of countries with emerging economies, such as Brazil, in addition to differences attributable to SES. The importance of cross-cultural studies has been well represented in several AL studies. A recent study scrutinizing the association between perceived stress and AL among Russians, Americans, and Taiwanese revealed important nation-specific sex differences in magnitudes and directions: strongest positive association for Russians of both sexes, weaker association among American women, and an unexpected negative association among Taiwanese men (Glei et al., 2013). The importance of urban and rural locations is also an important consideration, given a recent finding among Polish men who had lower AL levels if they were more educated, married, and residing in an urban area (Lipowicz, Szklarska, & Malina, 2013). In summary, it is imperative to expand international investigations among individuals living in different sociocultural contexts in order to understand unique vulnerabilities and culture-specific protective factors that propel pathways toward potential resilience. Resilient Hispanic Effects in Emerging Economies There is a paucity of research on resilience among inhabitants from emerging-economy

countries. Nevertheless, a burgeoning literature and several AL studies (for a review, see Beckie, 2012) report positive health outcomes among Hispanic populations, due perhaps to unique sociocultural contexts and value systems that might promote resilience (Gallo, Penedo, Espinosa de los Monteros, & Arguelles, 2009). Despite substantial adversities (e.g., immigration, discrimination, low incomes, less education, unequal opportunities, limited health care access), Hispanic populations living in the United States have at times shown equal and sometimes even better health outcomes in comparison to Caucasian populations (Gallo et al., 2009). Similarly, Hispanic individuals with non-American nativity exhibit better health than those born in the United States. Features of Hispanic culture (e.g., allocentrism, familyoriented values) represent attitudes and behaviors that might exert a protective influence on health outcomes (Gallo et al., 2009). The sociocultural dimensions of Brazilians exhibit similarities to those of Hispanics. Interestingly, Brazilian inhabitants score equally with populations from the developed world on psychometrics of resilience (Ferreira, Santos, & Maia, 2012; Fortes, Portugues, & Argimon, 2009; Gaiolli, Furegato, & Santos, 2012; Rodrigues, Barbosa, & Chiavone, 2013; Tavares, Barreto, Lodetti, Silva, & Lessmann, 2011) that actually increase with age (Gaiolli et al., 2012; Tavares et al., 2011). Among Brazilian children living under adverse conditions, a father's ability to adequately respond to his child's needs, ability to promote expression of the child's potential, and engagement in gratifying familial activities have been identified as key protective factors (Silva, Lacharité, Silva, Lunardi, & Lunardi-Filho, 2009). Similarly among young Brazilians from low-SES backgrounds, the cultivation of interpersonal relationships (e.g., family, friends, teachers) and personal resources (e.g., personality traits, optimism motivation, spirituality) promote resilience in the face of future life challenges (Silva et al., 2009). It is interesting to note that different studies reveal that resilience is independent of SES among adult and elder Brazilians (Gaiolli et al., 2012; Tavares et al., 2011). These protective factors include personality characteristics such as self-confidence, self-control, self-esteem, empathy, optimism, and sense of meaning (Rodrigues et al., 2013). Future crosscultural research could investigate these cultural distinctions by delineating important differences between the sexes around the world. As presented next, biological sex and sociocultural gender factors influence physical health and well-being.

Sex Differences and Gender Diversity Every cell is sexed, every person gendered, and every organism stressed (Juster & Lupien, 2012c). Whereas the term sex refers to a multidimensional construct that includes genes, anatomy, gonads, and hormones that collectively define us as male or female, gender refers to the implicit and explicit dissimilarities in an array of socioculturally constructed roles, orientations, and identities that generally predominate in one sex or the other (Juster et al., 2011). Relatedly, gender roles represent the social norms and stereotypes we assimilate and endorse through the enactment of masculine traits (e.g., independence, assertiveness, competitiveness, instrumentality) and feminine traits (e.g., empathy, compassion, helpfulness, expressivity (O. Evans & Steptoe, 2002; Johnson & Repta, 2011). One's sexuality also comprises its own constituent sexual behaviors, identities, and orientations that correspond

with health inequalities (Institute of Medicine [IOM], 2011). In the next sections, we summarize an ongoing research program on the effects of sex, gender roles, and sexual orientation on AL. Biological Sex and Aging Sex differences in AL depend on sample-based and age-related factors (Juster et al., 2010). For example, successive analyses of the MacArthur cohort provide evidence for age-specific sex differences in the biological signatures linking AL to tertiary outcomes, such as all-cause mortality. An early analysis found that cardiovascular biomarkers were more often dysregulated for men, whereas neuroendocrine biomarkers were more often dysregulated for women (Seeman, Singer, Ryff, Dienberg Love, & Levy-Storms, 2002). In a 12-year follow-up that used recursive partitioning, high-risk pathways of AL biomarker clustering for men included adrenaline, noradrenaline, IL-6, CRP, and fibrinogen, while women included IL-6, CRP, glycosylated hemoglobin, and systolic BP (Gruenewald, Seeman, Ryff, Karlamangla, & Singer, 2006). Interestingly, elevated systolic BP occurred in 100% of high-risk pathways for women but only 17% for high-risk pathways among men, principally driven by elevated fibrinogen, noradrenaline, and adrenaline levels otherwise absent for women. These findings from the MacArthur cohort measured at different time points reveal that sex differences in AL pathways might emerge only with advanced age. This is consistent with numerous studies demonstrating links between chronic stress and advanced aging at molecular, social, and environmental levels (Epel, 2009; Epel et al., 2004; Tomiyama et al., 2012). Identification of sex-specific thresholds in individual biomarkers and systemic clustering of AL at different ages should be a priority, given increasing older populations worldwide (Singer et al., 2004). In a first attempt, a 15-year follow-up of CARDIA participants in their 40s found no differences between sexes or by race/ethnicity. Nevertheless, it must be noted that aggregate metafactors including heart rate variability, BP, and inflammatory, metabolic, catecholaminergic, and glucocorcoticoid biomarkers captured 84% of the statistical variance, suggesting that AL biomarkers indeed form a constellation of shared interrelations consistent with theory (Seeman et al., 2010). Yet subtle physiological dysregulations of intertwined systems may not be easily detected or differ between the sexes when assessing younger adults. Consequentially, determining meaningful subclinical biomarkers cutoffs between sexes at different ages is essential. In this regard, sociocultural differences also must be considered cross-culturally. For example, a pilot study of Japanese elders measuring AL based on 33 parameters revealed that they were leaner with slightly more elevated BP than Europeans and Americans, in addition to having higher total HDL cholesterol (Crews et al., 2012). Despite a lack of age differences, sex differences were also found for dopamine, uric acid, glutamic-pyruvic transaminase (liver enzyme), and white blood cells. Although preliminary and requiring continued cross-cultural comparisons, we propose that delineation of sex- and age-specific differences will require gender-based perspectives that appreciate differential socialization patterns (e.g., social networks, diet and alcohol consumption, leisure activities) and the consequent risk and protection this engenders. In addition, do gender effects interact with sociocultural factors to

modulate AL in an international context? For instance and as discussed next, one could imagine the stressful circumstances of, say, a feminine Japanese American woman with conflicting traditions of collectivism versus individualism or perhaps a masculine male nurse, since each may embody mismatches to societal and occupational expectations of them. Sociocultural Gender Our understanding of sex differences can be complemented by incorporating gender perspectives when interpreting the AL literature (Juster & Lupien, 2012a). Sex differences in biological constitution, in concert with individual differences in elusive gender-based factors, are known to influence patterns of physiological stress responsivity (Dedovic, Wadiwalla, Engert, & Pruessner, 2009). For example, women generally self-report more environmental stressors and psychosocial distress, while men are generally more biologically stress responsive compared to women in the follicular phase of their menstrual cycle or on oral contraceptives. These sex differences stem from lifelong exposure to sex steroids, anthropometry, and sociocultural factors that construct one's gender that collectively drive diverse patterns of threat processing and coping mechanisms. It has been argued that sociocultural factors may be more important than biological factors in explaining mysterious sex differences in stress responsivity (Lundberg, 2005). For instance, it would seem that women demonstrate increased cortisol levels when facing social rejection (Stroud, Salovey, & Epel, 2002), whereas men tend to be more reactive when confronted with achievement-based stressors in laboratory paradigms (Dickerson & Kemeny, 2004). A complementary evolutionary hypothesis put forth by S. E. Taylor and colleagues (2000) states that men and women generally cope differently under stressful situations. Accordingly, the primary stress response pattern for men are fight-or-flight responses, whereas women might be more prone to engage in tend-and-befriend responses, such as nurturing and affiliationbased behaviors that protect against the demands of pregnancy, nursing, and child care (S. E. Taylor et al., 2000). The central tenet of Taylor et al.'s proposal is that instead of utilizing physical retaliation when faced with threats, women focus on behaviors that protect vulnerable offspring and maintain social bonds. Compelling animal and human studies have substantiated this theory; however, it is not clear whether these responses are due to sex, gender, or both—or perhaps even influenced by how our own gendered assumptions factor into our research designs (Juster & Lupien, 2012c). For instance, S. E. Taylor et al. underlined that prior to 1995, women represented a mere 17% of participants in human studies of stress due to the alleged complications of female reproductive functions. Importantly for the present discussion, the tend-and-befriend hypothesis sheds some light on sex and age differences found throughout the stress and AL literature vis-à-vis social networks and relational factors that likely also differ internationally according to socialization and gender stereotyping. For instance, AL levels were lower only for Taiwanese elderly men with a spouse, while both sexes benefit from ties with close friends and neighbors in a SEBAS analysis (Seeman et al., 2004b). Interestingly, the perceived quality of these social relationships was not consistently linked to AL, suggesting that cross-cultural differences to between Eastern and Western societies may be influenced by sociocultural factors (e.g., virtues

of collectivism/interdependence, social obligation, and the absence of “self”) that are inherently influenced by gender-based constructs related to societal patriarchy and/or matriarchy. In this spirit, one's sex and gender role might divergently influence biopsychosocial stress. In particular, extremely traditional or hegemonic masculine gender roles (e.g., hostile, dictatorial, impatient) are strongly linked to coronary-prone behaviors (e.g., inhibited emotional expression, lack of empathy, homophobia) that overlap with hostile personality traits shown to mediate the association between low SES with increased AL levels (Kubzansky, Kawachi, & Sparrow, 1999; J. Sun, Wang, Zhang, & Li, 2007). As women increasingly occupy nontraditional professions around the world (WHO, 2006), changing gender roles might incur both health-promoting and health-damaging consequences (Lundberg & Frankenhaeuser, 1999). It follows that individual differences in unique configurations of AL psychosocial antecedents might differ according to gender and related behaviors that might ultimately contribute to AL health consequences (Juster et al., 2011; Lupien et al., 2006). Our group recently investigated whether age, sex, and gender roles were associated with AL levels and physical complaints among Montreal workers from diverse occupations (Juster & Lupien, 2012b). Preliminary results revealed that: (1) increased masculinity (e.g., independent, assertive) and age predicted high AL levels, while sex did not; (2) increased masculinity and being female together predicted increased physical complaints; and finally, (3) high AL corresponded to increased physical complaints. That higher masculinity was related to increased objective physiological dysregulation and subjective physical complaints (e.g., nausea, dizziness, and pain) suggests an overall increased vulnerability to hyperarousal pathologies such as cardiovascular disease among masculine-typed individuals. These triangulated findings are clinically important. In particular, atypical prodromal symptoms of acute coronary syndrome for women include physical complaints (Pilote et al., 2007). This fact provides preliminary support for the inclusion of gender roles in studies investigating AL with differences otherwise attributable solely to sex or age. Interpreting sex differences in pathophysiology and AL levels involves consideration of gender-based factors that need to merge qualitative and quantitative approaches. For example, widespread assumptions that men are vascular reactors and women that are cardiac reactors do not address intra-individual differences between, as well as within, the sexes (Legato, 2010). Many of the antecedents of AL, such as social support, personality traits, occupational characteristics, and lifestyle behaviors (see Figure 8.1), are influenced by sociocultural factors for which gender perspectives can be applied. In synthesizing existing literature with future directions in mind, we postulate that the synergism or interactions among sex- and genderbased factors modulate biological signatures (diurnal cortisol, reactive cortisol, AL) more strongly than the main effects of either set of factors alone (Juster et al., 2013a). In addition to sex and gender roles, one's sexual orientation has received limited attention in the biopsychosocial stress and AL literature, an area to which we turn to next. Sexual Orientation

Lesbian, gay, and bisexual (LGB) individuals, as a result of their minority sexual identities, frequently experience stress resulting from living in a society that negates or denigrates their identity. Specifically, LGB individuals face heterosexism on a regular basis. Herek (1992) defined heterosexism as an ideological system that denies, denigrates, and stigmatizes any nonheterosexual form of behavior, identity, relationship or community. Heterosexism encompasses homophobia, an irrational fear or hatred of nonheterosexual individuals, as well as bias against LGB individuals. Manifestations of heterosexism can be subtle, such as assuming someone is heterosexual, or overt, such as using anti-LGB epithets, engaging in violence against LGB individuals, or discriminating against LGB people in hiring or housing. Experiences of heterosexism are common in the lives of LGB individuals. For example, Mays and Cochran (2001) found that LGB individuals were twice as likely as heterosexuals to have experienced prejudice. Likewise, Herek, Gillis, and Cogan (1999) found that up to 25% of their LGB participants had experienced physical violence or property crimes because of their sexual orientation. Recent national data from the Canadian Centre for Justice Statistics show that hate crimes based on sexual orientation are the third most prevalent type of hate crime after race/ethnicity and religion. Moreover, hate crimes based on sexual orientation are characterized by violence more often than other types of hate crimes. For example, 56% of hate crimes based on sexual orientation in 2006 were violent, whereas 38% and 26% of hate crimes based on race/ethnicity and religion, respectively, were violent (Dauvergne, Scrim, & Brennan, 2008). Both overt and subtle heterosexism increase stress for LGB individuals. This increased stress can lead to AL and negative health outcomes. This form of sociocultural stress has been termed minority stress (Meyer, 1995; Meyer, 2003), which refers to the stress of living in an environment that is stigmatizing of nonheterosexual sexual orientations. Meyer (2007) identified three overarching characteristics of minority stress: 1. Minority stress is unique. The stress of living in an environment that stigmatizes nonheterosexual sexual orientation is unique to LGB individuals and different from mundane stressors encountered by all people and moreover is additive to the stress caused by normal everyday stressors. 2. Minority stress is chronic. LGB individuals experience this form of stress on a daily basis from events ranging from mundane offenses to extreme instances of harassment and violence. 3. Minority stress is socially based. Minority stress is caused by factors outside the individual by other people, institutions, and social-political processes. Meyer (1995, 2003) divided minority stress into five separate components: (1) internalized homophobia, (2) perceived stigma, (3) prejudice events, (4) sexual orientation concealment, and (5) subsequent coping strategies. Internalized homophobia refers to negative views of homosexuality/bisexuality that have been internalized by the LGB individual. Perceived stigma relates to the view that a person will be treated unfairly because of her/his sexual orientation. Prejudice events are the actions of others toward a LGB individual that are discriminatory,

biased, or violent; these can vary in intensity from overt physical or verbal attacks to more subtle instances of bias. Sexual orientation concealment refers to attempts to hide one's sexual orientation or “pass” as heterosexual. These components of minority stress require individuals to engage in compensatory coping strategies in an attempt to ameliorate negative effects. Minority stress is linked to negative health outcomes. Indeed, a recent review of the literature by the Institute of Medicine (2011) highlighted that LGB individuals tend to have elevated levels of psychological distress, increased incidence of suicidal ideation and attempts, and increased substance use relative to their heterosexual peers. Early studies noted the high incidence of minority stressors and their relationship to health outcomes. For example, Ross (1990) found that the majority of his sample of gay men had heard antigay jokes, had to associate with someone who was homophobic, and had been harassed or threatened. In addition, 25% of his sample had been the victim of antigay physical assault. Moreover, these experiences of discrimination were positively related to somatic symptoms, anxiety, and insomnia. Other studies have continued to demonstrate links between minority stress and a variety of negative mental, physical, and behavioral health outcomes. As reported by the Institute of Medicine (2011), the negative outcomes of minority stress include psychological distress, anxiety, depression, suicidal behaviors, insomnia, physical symptoms, health complaints, substance abuse, prostitution, increased number of sexual partners, and unprotected anal intercourse to name but a few. In particular, behavioral outcomes associated with minority stress include substance use and unprotected sexual intercourse that are important risk factors for the transmission of the human immunodeficiency virus (HIV) among gay and bisexual men in particular. The high incidence of HIV among gay and bisexual men—and the role of minority stress as a risk factor for HIV transmission—has been explained by the concept of syndemics. A syndemic is defined as “a set of closely intertwined and mutual enhancing health problems that significantly affect the overall health status of a population within the context of a perpetuating configuration of noxious social conditions” (Singer, 1996, p. 99).With regard to gay and bisexual men, psychosocial health problems, including depression, substance use, and antigay victimization, are related to HIV risk. When these psychosocial variables combine, the resulting impact on HIV risk is greater than the sum of their individual effects (Safren, Reisner, Herrick, Mimiaga, & Stall, 2010; Stall, Friedman, & Catania, 2008; Stall et al., 2003). For example, Mustanski, Garofalo, Herrick, and Donenberg (2007) found that gay and bisexual men who engaged in binge drinking were over five times more likely to use street drugs; those who used street drugs were almost four times as likely to engage in unprotected anal intercourse. Moreover, they found that for each additional psychosocial health problem (e.g., alcohol use, drug use, psychological distress, violence), the odds of engaging in unprotected anal intercourse increased by 42%. Research on syndemics suggests gay and bisexual men's increased HIV risk can be understood in the context of living in a stigmatizing environment that results in victimization, mental health problems, and substance use. From a transdisciplinary perspective, these sociocultural outcomes interact with each other syndemically to increase risk for HIV.

The accumulated findings over the past several decades suggest that minority stress plays an important role in the psychological, physical, and behavioral health of LGB individuals. However, this rich body of psychological research is matched by a paucity of biopsychosocial research. Research examining biological data among LGB individuals has been sparse, with earlier research examining biological outcomes among HIV-positive gay and bisexual men (and sometimes matched HIV-negative controls), and research focusing on healthy LGB populations only emerging later. Most of the research that includes a focus on biological data has examined the relationship between psychological distress and physical health, often in HIV-positive gay and bisexual men. For instance, Leserman and colleagues (2000) found that faster progression to AIDS among HIV-positive gay and bisexual men was associated with more stressful life events, use of denial coping strategies, lower satisfaction with received social support, and lower serum cortisol levels. Greeson et al. (2008) found that the relationship between HIV disease progression and psychological distress was mediated by decreased natural killer cell immunity (fewer cells and less cytotoxic function) and increased cytotoxic T-cell function. In a study of both HIV-positive and HIV-negative gay men, Kertzner Goetz, Todak, and Cooper (1993) found that urinary free cortisol was related to depressed and anxious moods. Resilience When “Coming Out of the Closet” Early research on the minority stress factor of sexual orientation concealment revealed that both HIV-positive and HIV-negative gay men evidenced worse health outcomes when they concealed their sexual orientation. Cole and colleagues (1996b) found that HIV-negative gay men who were closeted had significantly higher incidence of cancer and infectious diseases than those who were out. In addition, HIV-positive gay men who were closeted reached a critically low CD4 cell count, received an AIDS diagnosis, and died significantly faster than those who were out (Cole, Kemeny, Taylor, Visscher, & Fahey, 1996a). Likewise, among healthy LGB, disclosure represents an important source of distress. Indeed, “coming out of the closet” (Riley, 2010) to family, friends, and coworkers is considered one of the most significant stressors in the lives of LGBs (Hershberger, Pilkington & D'Augelli, 1997). To date, only one study of healthy sexual minority men showed that disclosure at work was unexpectedly associated with higher cortisol levels and negative affect in comparison to nondisclosed workers (Huebner & Davis, 2005). Shedding light on this unintuitive finding are those from a qualitative study showing that gay men who disclosed their sexual orientation to supervisors reported significantly more hostility in their work environment, significantly fewer perceived promotion opportunities, and significantly higher turnover intentions (Tejeda, 2006). Disclosure of one's sexual orientation is a complex phenomenon that has been purported to involuntarily incite harassment and discrimination among work colleagues (Waldo, 1999) and in the community more broadly (Huebner, Rebchook, & Kegeles, 2004). As noted, however, disclosure can eventually promote positive health and well-being, as evidenced by less distress reported than with nondisclosure (Greene, Derlega, & Matthews, 2006; Morris, Waldo & Rothblum, 2001). A recent study by our group assessed psychiatric symptoms, diurnal cortisol, and AL levels as

a function of sexual orientation as well as disclosure status. The results revealed that sexual minorities did not manifest more stress-related problems than heterosexuals (Juster, Smith, Ouellet, Sindi, & Lupien, 2013b). Quite to the contrary, gay/bisexual men had lower depressive symptoms and AL levels than heterosexual men, which is inconsistent with a sexual minority stress model. Differences in AL levels were driven by lower levels of triglycerides, body mass indices, and TNF-α levels among the gay/bisexual men that could be explained behaviorally by their increased focus on thinness and muscularity by means of diet and exercise compared to heterosexual men (Juster et al., 2013b). Alternatively, the differences may be due to sample-based features or potentially even be influenced by resilient pathways related to the rites of passage associated with sexual minority status and coming out. Given that two-thirds of LGB participants had completely disclosed their sexual orientation, we ascertained whether disclosure might explain these surprising reversals. Accordingly, those who had completely come out to friends and family had less symptoms of anxiety, depression, and burnout as well as lower levels of morning cortisol than those who remained closeted (Juster et al., 2013b). In a secondary analysis by our group, retrospective coping strategies used during disclosure (problem-focused, seeking social support, wishful thinking, blaming self, and avoidance) were assessed in relation to current indices of environmental, psychological, and biological stress. This period of sexual identity formation represents the elapsed time since first recognizing same-sex attraction or disclosing to family and friends. The increased use of avoidance coping behaviors was associated with increased self-reports of daily hassles and perceived stress, as well as increased AL levels. Interestingly, coping that involved seeking social support buffered against current perceived stress. Could prosocial behaviors and disclosure represent respective coping strategies and life transitions that promote resilience among sexual minorities? Stigma-related stress can promote individual differences in adaptive behavioral responses among stigmatized individuals that can render them more resilient and effective at managing future stressors, especially if they are well supported socially and are able to develop adaptive coping strategies (Hatzenbuehler, 2009). Although these novel findings link sexual orientation, life transitions, and coping strategies to psychological outcomes, cortisol, and/or AL levels, we cannot assume that our results generalize to all LGB communities. Specifically, participants were all recruited from the liberal city of Montreal, and Canadian social policies are progressive vis-à-vis sexual minority rights. As such, our findings might not generalize to more conservative locations in particular. Moreover, differences among specific LGB groups intersect according to age cohorts, race/ethnicity, SES, culture, and geographic location (IOM, 2011). For instance, younger generations of LGBs are 16 times more likely to have disclosed than older generations (Gates, 2010), and African American gay/bisexual men experience higher levels of internalized homophobia and are therefore less likely to disclose than other racial/ethnic groups (Peterson & Jones, 2009). Intersectionality is an inductive research framework that appreciates the reality that individuals are members of multiple social groups and identities that will contextually determine their experiences, chances, opportunities, and ultimately health (Johnson, Repta & Kaylan, 2011).

Moreover, there are many environments where disclosure may be a dangerous option. At a societal level, however, these findings have social implications: If individuals are to garner any benefits from disclosing, there must be established social policies that reduce stigma and intolerance (Chaudoir & Fisher, 2010). Institutionalized stigma and heterosexism includes, for example, the denial of marriage rights, disadvantaged treatment in schools and workplaces, and disenfranchisement of sociocultural resources like religion and spirituality that often dehumanize LGBs and contribute to psychosocial distress (Herek & Garnets, 2007; IOM, 2011). These social inequalities have health consequences that can be reversed when policy changes are made (Hatzenbuehler, 2010). Compelling research shows that American LGBs living in states without policies that protect them from hate crimes and employment discrimination experience significantly higher rates of psychopathology than LGBs living in states with protective policies (Hatzenbuehler, Keyes & Hasin, 2009). Likewise, LGBs living in states with constitutional amendments banning same-sex marriages experience increased rates of generalized anxiety disorder, depressive disorders, and alcohol abuse. By contrast, those living in states that recognize same-sex marriages show no increased development of these conditions (Hatzenbuehler, McLaughlin, Keyes, & Hasin, 2010). A recent study showed that following the Massachusetts legalization of same-sex marriage, there were significant decreases in general medical and mental health care visits and costs 12 months afterward among gay men (Hatzenbuehler, O'Cleirigh, Mayer, Safren & Bradford, 2012). This body of research demonstrates how changes in distal policies can progressively eliminate institutionalized stigma and promote public health benefits (Hatzenbuehler et al., 2012). In a pioneering study, Hatzenbuehler and McLaughlin (2013) examined stress-reactive HPA axis functioning emphasizing discrimination among sexual minorities exposed to varying degrees of stigma. Results indicated that LGB young adults who lived in states with high structural stigma (e.g., states without policies that protect sexual minorities from hate crimes, ensure employment nondiscrimination, and support same-sex marriage) as adolescents showed blunted cortisol responses following the Trier Social Stress Test (TSST) compared to those from low-stigma states. This finding suggests that the stress of growing up in environments that target sexual minorities for social exclusion may result in HPA axis dysfunction (Hatzenbuehler et al., 2014). Given ongoing debate in the United States and worldwide concerning same-sex marriage, assessing AL levels in states with and without protective policies over time would contribute further to our understanding of the relation between social policy and biology. Beyond sex and gender diversity, the plight of Indigenous peoples within our national borders must not be ignored or forgotten. Especially troubling is the fact that few biopsychosocial studies have been conducted among Indigenous peoples, especially for studies quantifying chronic stress. In the next section, we discuss the American Samoans, Nunavik Inuit, and Australian Aborigines as Indigeneous populations at risk of higher AL.

Indigenous Peoples of North America and Australia American Samoans

American Samoa is a group of five volcanic islands and two coral atolls located in the South Pacific that is an unincorporated U.S. territory (Central Intelligence Agency, 2012). As early as 1000 B.C., proto-Polynesians settled the islands. Today, native American Samoans are descendants of the Lapita culture and are one of the few remaining Polynesian societies (Bellwood, 1987). In 2012, the native American Samoan population was estimated at 49,782 individuals (Central Intelligence Agency, 2012). American Samoan women have a life expectancy of 75.9 years in comparison to 69.3 years for men (WHO, 2011), which is about 5 to 7 years lower than that of the general American population as of 2009 (80.9 years for women in comparison to 76.0 for men [WHO, 2012]). Migration of many Samoans to Hawaii and the American mainland has led to a number of communities with widely varying sociocultural characteristics. Investigations on the challenges of changing social contexts and acculturation on health have been undertaken (Baker, Hanna, & Baker, 1986). Public health reports reveal that American Samoans have alarming rates of metabolic disorders. American Samoans have the highest prevalence of adulthood obesity in the world: 74.6% of the adult population classify as obese, according to their BMI (WHO, 2011). Additionally, diabetes mellitus cases have doubled and are a leading cause of death for American Samoans (WHO, 2011). Metabolic diseases are strongly associated with chronic stress (Kyrou & Tsigos, 2007, 2009; Mattei, Noel, & Tucker, 2011; Stumvoll, Tataranni, & Bogardus, 2004; Tamashiro, Sakai, Shively, Karatsoreos, & Reagan, 2011) that can synergize with acculturated dietary factors to increase AL and therefore increase vulnerability to diabetes mellitus. In a landmark study, it was shown that Pima Indians living traditionally in rural Mexico were in better health than those living in Arizona for whom diabetes incidence was much higher (Ravussin, Valencia, Esparza, Bennett, & Schulz, 1994). Studies that can assess such adverse acculturation and consequent AL could provide powerful evidence to inform policymakers. In a first study, Crews (2007) investigated the association between AL and diabetes mellitus among American Samoans. AL was significantly associated with diabetes mellitus (Crews, 2007), consistent with higher prevalence among American Samoans. While Samoan inhabitants are well integrated into a cosmopolitan way of life, recent acculturation has led to increasingly overlapping sociocultural risk factors for metabolic diseases common to the larger American population. As a consequence, clusters of “new” risk factors (e.g., industrial food production) are generally lower in the oldest portion of the population (Bindon, Knight, Dressler, & Crews, 1997) who grew up and developed before these massive changes in their lifestyle. Interestingly, AL was negatively associated with age in men but was more strongly positively associated with age in women. In sum, the oldest Samoans are those who grew and developed before massive sociocultural changes, the middle-age Samoans experienced many of these changes, and the youngest generation are the now the most obese Samoans (Crews, 2007). These differences may be driven by interactions among age, biological sex, and sociocultural gender-based differences like those discussed in the previous section. It is important to highlight that gender issues are of high importance in Samoan society, with active promotion and protection of women's rights and equalities in addition to equality with men with respect to participation in the paid labor force, educational attainment, and occupation of several senior

positions in the public sector (WHO, 2011). Yet sociocultural changes may have caused greater disruption of women's traditional way of life that may have contributed to their higher levels of AL. Although the exact mechanism is unclear, obesity, hypertension, and diabetes are more common among older Samoan women in contrast to men (Baker et al., 1986; Crews, 1989). This suggests that AL might reflect sociocultural changes in the life style of Indigenous populations and might be useful to predict future health vulnerabilities to metabolic disorders among others. Nunavik Inuit The Inuit are a group of native inhabitants of the Arctic regions of Russia, the United States, Canada, and Greenland that share similar culture and ancestors as the descendants of the Thule culture that migrated from eastern Asia to Alaska by crossing the Bering land bridge around 1000 BC. By the thirteenth century, the Inuit had settled across the continent toward eastern Arctic territories of Canada, such as Nunavik (McGhee, 1978). Nunavik is a vast territory (larger than California) north of the fifty-fifth parallel in the province of Quebec, Canada. In 2006, the Nunavik Inuit population comprised 10,950 people living in 14 coastal communities (Statistics Canada, 2006a). With a median age of 20.8 years, they have the lowest mean age in Canada (Statistics Canada, 2006a). Nunavik Inuit are the most vulnerable Canadian Inuit population to health problems (Statistics Canada, 2006b, 2006c, 2008). Since 2004 to 2008, they have had the lowest life expectancy at birth in Canada and the lowest overall life expectancy, at 66.5 years (68.5 years for females and 64.5 years for males) in comparison to 80.6 years for the Canadian average (82.9 years for females and 78.2 years for males (Statistics Canada, 2008). Moreover, life expectancy is 5.6 years lower than the average of any other Canadian Inuit (Statistics Canada, 2008). The AL model could be applied to further understand the social determinants of these disturbing demographics. In particular, housing conditions are a major source of distress (G. W. Evans, Lepore, Shejwal, & Palsane, 1998) and a potential determinant of compromised Indigenous health (Gracey & King, 2009). Approximately half of Nunavik Inuit live in overcrowded houses, defined as more than one person per room, compared with 33% for the average of any other Canadian Inuit and 3% for non-Indigenous Canadian households (Statistics Canada, 2006c). This situation is explained in part by high birth rate, housing shortage, and high cost of construction (Rochette, St-Laurent, & Plaziac, 2007). The association between household crowding and AL was recently assessed by our group using data from the Nunavik Inuit Health Survey (Blanchet & Rochette, 2008; DeWailly et al., 2007a, 2007b; Kirmayer & Paul, 2007; Rochette et al., 2007) that calculated AL by incorporating 14 biomarkers from on a cross-section of 839 Inuit between the ages of 18 and 79. Results show that higher household crowding was significantly associated with elevated AL levels while controlling for the effects of age, sex, income, diet, and involvement in traditional activities (Riva et al., 2014) Another major risk factor for the Nunavik Inuit that overlaps with household conditions is low SES. In particular, educational level is very low in Nunavik: Only 25% of Nunavik Inuit over

15 years old have a high school diploma or above (Statistics Canada, 2006b). Unemployment is also an important issue for the Inuit from Nunavik with only 56.1% over 15 years of age employed (Statistics Canada, 2006b). The proportion of this population with stable and wellpaying employment is also relatively limited in Nunavik: income is exceedingly low (42.6% earning less than $20,000/year) and only 32.2% have full-time employment (Statistics Canada, 2006b). This low income in turn leads to food scarcity reported by 25% Nunavik Inuit (Blanchet & Rochette, 2008). Finally, exposure to environmental contaminants is a well-known problem in the Canadian Arctic. Members of our group have previously proposed that environmental toxins might modulate AL and exacerbate health outcomes (Juster et al., 2011; Loucks, Juster, & Pruessner, 2008; McEwen & Tucker, 2011). Accordingly, a recent cross-sectional analysis of the NHANES cohort demonstrated that high AL levels amplify the damaging effects of exposure to lead in terms of elevated BP (Zota, Shenassa, & Morello-Frosch, 2013). This is likely much more problematic for Nunavik Inuit inhabitants who are exposed to a plethora of environmental contaminants that are carried from southern to northern latitudes by atmospheric and oceanic currents that are then bioaccumulated in the Arctic food chain. Specifically, the traditional Inuit diet contains high levels of fish (Arctic char) as well as land (caribou) and marine mammals (beluga, seal, and walrus) that are at the top of the food chain (Blanchet & Rochette, 2008). Consequently, the Nunavik Inuit are more exposed to environmental contaminants than the average population from southern Quebec (Muckle, Ayotte, Dewailly, Jacobson & Jacobson, 2001). This exposure includes exposure to traditional legacy contaminants (heavy metals like lead, methylmercury, and cadmium) and persistent organic pollutants (mainly polychlorinated biphenyls [PCBs]) and chlorinated pesticides (DeWailly et al., 2007a, 2007b). Environmental contaminants can exert endocrine-disrupting effects that modulate allostatic processes (Juster et al., 2011). For example, among wildlife, impaired ability to mobilize glucocorticoids in response to an acute stress is related to heavy metal exposure and environmental organic contaminants (e.g., PCBs, solvents, pesticides) in fish (Benguira, Leblond, Weber, & Hontela, 2002; Hontela, Dumont, Duclos, & Fortin, 1995; Hontela, Rasmussen, Audet, & Chevalier, 1992), amphibians (Gendron, Bishop, Fortin & Hontela, 1997), birds (Love et al., 2003), and large mammals (Oskam et al., 2004). In rodents, altered HPA axis functioning has been reported following early exposure to PCBs (Meserve, Murray, & Landis, 1992) and heavy metals (Vyskocil, Fiala, Ettlerova, & Tenjnorova, 1990). Human studies demonstrate that some PCBs even act as antagonists of glucocorticoid receptors (Johansson, Nilsson, & Lund, 1998) and alter glucocorticoid biosynthesis in human adrenocortical cells (L. A. Li & Wang, 2005). This finding has also been shown at low levels of lead exposure: Exposure to lead increases glucocorticoid responses to acute stress in 9- to 10-year-old children (Gump et al., 2008). Likewise, a recent study reported an association between lead exposure and higher cortisol levels in occupationally exposed workers (Fortin et al., 2012). Beyond the HPA axis, environmental toxins influence cardiovascular functioning. For example, elevated exposure to methylmercury is pathophysiological and increases heart rate, BP, and coronary heart disease (Ayotte et al., 2011; Valera, Dewailly, & Poirier, 2008, 2009, 2013; Valera et al., 2012). Finally, prenatal exposure to PCBs has been related to

incidence of acute respiratory infections in Inuit children (Dallaire et al., 2006). Numerous environmental contaminants are related to adverse mental health outcomes. For Inuit infants, postnatal lead exposure is linked to increased inattention, impulsivity, and irritability (Plusquellec et al., 2007, 2010), while pre- and postnatal PCB exposure is linked to increased unhappiness, anxiety, and inattention (Plusquellec et al., 2010; Verner et al., 2010). This finding is clinically important as postnatal lead exposure increases the risk of hyperactivity and attention-deficit disorder in Inuit children (Boucher et al., 2012), while prenatal methylmercury exposure increases inattention in Inuit infants (Verner et al., 2010). Thus, environmental contaminant exposure may alter stress physiology and influence developmental psychopathology. In summary, comprehensive biopsychosocial studies of Indigenous people in North America are needed. As reviewed, American Samoans and Inuit from Nunavik are both at high risk of chronic stress through socioeconomic factors, health behaviors, and environmental contaminants. There is therefore a strong need for continued research using the AL model to understand the health disparities of Indigenous peoples and to inform public health policies. Turning now to Oceania, we explore how the AL model can be further cross-culturally tailored to the challenges of Australian Indigenous peoples. Australian Aborigines Aboriginal people have lived in Australia for at least 60,000 years. Today, there are two main groups of Indigenous Australians: those from the Australian continent and the island state of Tasmania and those from the Torres Strait. With an estimated population of 575,552 and projected annual growth rate of 2.2%, the Aboriginal and Torres Strait Islander population of Australia represent 2.5% of the total Australian population. The Australian Aboriginal population is relatively young, with a median age of 21 years compared to 37 years for the non-Indigenous population as of 2006 (Australian Bureau of Statistics, 2007). Prior to the arrival of European settlers in the late eighteenth century, Indigenous Australians were healthier than most Europeans (Webb, 2009). As elsewhere around the world (e.g., among Native American communities), European settlers arrived with diseases that subjected Indigenous people to widespread changes that included loss of land and livelihood, as well as restricted civil rights and forced familial breakdown by religious missions that removed Indigenous children from their families (Chesterman & Galligan, 1997). These dramatic changes have destroyed their traditional way of life and have had abysmal effects on their health and well-being. Mortality rates are much higher among the Australian Indigenous population when compared to the general Australian population. Causes of morbidity and mortality include cardiovascular disease, diabetes, and tobacco-related conditions that collectively account for half of health disparities in comparison to non-Indigenous Australians (T. Vos, Barker, Begg, Stanley, & Lopez, 2009). Life expectancy is 67.2 years for men (11.5 years lower than for non-Indigenous men) and 72.9 years for women (9.7 years lower than for non-Indigenous women). Roughly 87% of all Australians reach age 65; this is reduced to 24% and 35% of Australian Aboriginal

men and women, respectively. These statistics are lower than in Nigeria, Bangladesh, and several other developing countries (UN, 2003). Cardiovascular diseases and diabetes are much more common for Indigenous people than for non-Indigenous people (Australian Bureau of Statistics, 2006) and have a much younger incidence. Young and middle-aged Indigenous adults surprisingly drive these differences: the death rates for the 35 to 44 age groups and 45 to 54 year age groups are 8 to 12 times higher than corresponding non-Indigenous rates by age (Australian Institute of Health and Welfare, 2008). Although the prevalence of diabetes increases with age, here too the pattern of onset is much earlier: prevalence for Indigenous people aged 25 to 34 years old was almost seven times that of non-Indigenous people, while the prevalences reported by Indigenous people aged 35 to 44 as well as 45 to 54 years are more than five times those reported by age-matched nonIndigenous people (Australian Bureau of Statistics, 2006). This rate represents a strain on the health care system, as hospitalization rates for diabetes among Indigenous men and women are 3.4 and 5 times, respectively, the rates of non-Aboriginals after adjusting for age (Australian Institute of Health and Welfare, 2011). Chronic stress and AL likely mediate associations among social inequalities and health disparities among Australia's Indigenous population. Like previous cross-cultural AL studies (e.g., Taiwanese SEBAS cohort), psychosocial stressors—such as death of a family member or closed friend, serious injury or disability—are major contributors of cardiovascular and metabolic problems among Australian Indigenous people (Australian Institute of Health and Welfare, 2008). In a large community sample of Australian Aboriginal and Torres Strait Islanders, glycosolated hemoglobin was measured and interpreted as a biological indicator of social environmental stress (Daniel, O'Dea, Rowley, McDermott, & Kelly, 1999). As an important biomarker of diabetes, mean glycosylated hemoglobin concentrations were 18.2% higher for both Indigenous groups in comparison to Caucasian Australians and for Greek migrants who experience unique distress due to acculturation. Despite recent major efforts to study the social determinants of disease that disproportionately affect Indigenous Australians, relatively few studies have investigated AL biomarkers in community samples to serve as contrasts. Schmitt and colleagues (1995) studied urinary cortisol and adrenaline excretion rates in three Aboriginal communities in the Kimberley region of northwest Australia and compared them to samples from Oxford in the United Kingdom. They found that both cortisol and adrenaline secretion was significantly higher in the Aboriginal communities than in the Oxford sample, which was already quite high by international standards. Mental health burden is similarly much higher among Indigenous people. Psychiatry-related hospitalizations have also increased: schizophrenia, schizotypal personality disorder, and delusional disorders are 4 times higher, substance use disorder is 3.8 times higher, and suicide attempts are 2.7 times higher among Indigenous people. High rates of psychosis and alcohol and cannabis use have also been reported among the Indigenous populations of Cape York and the Torres Strait, who experience heightened distress (Hunter, 2013; Hunter et al., 2012). Large-scale surveys reveal that prevalence for symptoms of depression and anxiety are 3 times

higher than for non-Indigenous adults (Jorm, Bourchier, Cvetkovski, & Stewart, 2012). Qualitative research among Aboriginal men in central Australia reveals that depressive symptoms are mostly attributable to a sense of loss and connection with the social and cultural fabric of Aboriginal life as well as with the cumulative stress of marginalization (Brown et al., 2012). Low SES (Cunningham & Paradies, 2012) and other social inequalities, including transgenerational trauma, discrimination, and racism, may underlie these health inequalities among Indigenous Australians (Marmot, 2011). Research that can further assesses intersections within marginalized groups (e.g., sexual minority Indigenous peoples) are needed. These statistics demonstrate pernicious effects of social marginalization among Indigenous populations that ultimately exacerbate trajectories toward mental and physical morbidity. Utilizing the AL model to demonstrate this conclusively would provide an important platform by which to assess the needs of marginalized groups and in turn inform policymakers of intervention efficacy that address these needs. In conclusion to this first part dedicated to social inequalities, we have reviewed an impressive breadth of research linking AL to mental and physical health. Studies clearly demonstrate that lower SES is associated with greater risk of increased AL, concomitant with important inequalities in Indigenous status, race/ethnicity, geography, as well as differences related to sex, gender, and sexual orientation throughout life that ought to be addressed in the evolution of public policies. Fundamental conclusions surface from this analysis of AL. AL represents a reliable measure of physiological dysregulations underlying illnesses that are influenced by social inequalities, such as with socioeconomic gradients, discrimination, homophobia, and adverse environments. In the interest of promoting continued advances in the measurement of AL in relation to biological, neurological, and cognitive indices of vulnerability and resilience to pathologies, we turn next to innovative approaches that should be incorporated into future research endeavors.

Innovative Biochemical, Neurological, and Cogntive Perspectives In the first part of this chapter, we explored the psychosocial aspects of social inequalities influencing vulnerability toward, and resilience against, AL. Elevated AL hastens the development of age-related disease outcomes, such as cognitive and functional decline, cardiovascular disease, and mortality. However, the cellular mechanisms by which social inequalities and adverse experiences promote disease remain elusive. More specifically, the “primary effects” of allostatic mediators in the AL model (e.g., specific alterations in cellular functioning) have received limited empirical substantiation. The reader is reminded that primary mediators (glucocorticoids, catecholamines, cytokines) exact a domino effect on secondary outcomes (e.g., glycol-lipid profiles, cardiovascular functioning) via primary effects—that is, cellular changes in tissue and organ functioning. Allostatic mechanisms like the stress response are diverse and span the spectrum of adaptive and maladaptive response patterns. In considering stress and resilience at a biological level—

or what is referred to herein as biological resilience—we wish to first define three key interconnected concepts outlined by Karatsoreos and McEwen (2011) regarding resilience, resistance, and recovery. First and consistent with previous sections, resilience is defined as an organism's ability to rebound from adversity when its ability to physiologically function has been tampered with in some negative way; however, the organism is able to adapt by activating allostatic mechanisms that can promote resistance. Second, resistance is defined as an organism's ability to withstand adversity and face future stressors with little or no stress response. Resistance is conceptually akin to vaccination or stress inoculation that helps bolster immunity to future stressors. Third, recovery is defined as an organism's ability to shut down the stress response and other related biological activities back to baseline levels (Karatsoreos & McEwen, 2011). All three of these constructs are related to adaptation at a biological level and are central to our understanding of biological processes that promote resilience. Although the AL framework is embedded within such processes, we present diverse research areas to provide food for thought to help guide future research. In this section, we focus on biological pathways that might mediate the adverse effects of chronic stress on health and ultimately modify vulnerability and resilience across the life span. We begin this discussion at a cellular level central to energy metabolism and the aging process —namely, the mitochondrion (Picard, 2011). Next, we present advances in hippocampal neurogenesis to underline an important development in neurosciences that has redefined our understanding of neural plasticity of the hippocampus. Finally, we explore the concept of cognitive reserve among the elderly to highlight continued research on AL and cognitive functioning. In the next three sections, we highlight the importance of assimilating new techniques and perspectives to refine our understanding of social inequalities in AL and navigating potential routes toward resilience.

Mitochondria Each of the approximately 10 trillion cells of the human body contains hundreds of mitochondria within their cytoplasm. Mitochondria are cellular organelles that arose 2.2 to 2.5 billion years ago when the ancestor of the mammalian cell fused or engulfed an aerobic bacteria capable of utilizing oxygen for energy production (Sagan, 1967). The bacteria thereafter remained and evolved into today's mitochondria. Through this endosymbiotic fusion, the mammalian cell de novo acquired from the ancient bacterium—now mitochondria—the capacity to produce large amounts of energy by oxidative phosphorylation (OXPHOS), which allowed the evolution of complex genomes and multicellular life-forms (Lane & Martin, 2010; Wallace, 2010). Cells of the human brain, muscles, vasculature, heart, and other organs, therefore, rely heavily on mitochondria for energy production and other cellular functions that sustain life (Scheffler, 1999). For this reason, mitochondria that do not function normally can cause death in infancy or severe multisystemic diseases in childhood and adulthood (Koopman, Willems & Smeitink, 2012). Beyond regulating life and death at the cellular level, mitochondria also produce signals within cells (Liu & Butow, 2006), between cells (Taivassalo, Ayyad, & Haller, 2012), and across organ systems (Taivassalo et al., 2003) that shape the body's allostatic mechanisms and thus the capacity to adapt to environmental

challenges. As reviewed in previous sections, environmental and perceived stress initiates allostatic mechanisms and cellular processes. These adaptive physiological processes involve metabolically expensive neuronal activities, increases in heart rate and BP, secretion of neurochemicals, and hormone synthesis. At a more fundamental level, they require changes in gene expression regulated in part by epigenetic writers and modifiers that ultimately regulate protein synthesis. All of these adaptive processes are energy dependent and thus increase cellular energy demand. Given that mitochondrial function provides the necessary energy reserve to fuel these adaptive processes, mitochondria have been regarded as a key component of the stress response (Manoli et al., 2007). These organelles also contain receptors for glucocorticoids (S. R. Lee et al., 2013) as well as other steroid hormones (Psarra & Sekeris, 2009; Yang et al., 2004) that regulate mitochondrial function in the brain and throughout the body (Rettberg et al.,2014). Taken together, this evidence attests to the importance of conceptually linking the biological stress response to mitochondrial function. Differences in mitochondrial content and function that determine cellular energy capacity could therefore contribute to vulnerability or resilience to stress-related diseases (Morava & Kozicz, 2013). For example, following acute stress, the combination of a rapid rise in cellular energy demand with insufficient endogenous energy capacity to sustain this demand could cause a bioenergetic mismatch. Because cellular energy is a rate-limiting step for many protective cellular processes, a transient mismatch causing energy deficiency would generate cell stress, damage, and possibly elevated AL over time. Applied to stress physiology, inadequate capacity for mitochondrial energy production may thus be a source of cellular AL. Conversely, fostering a reserve capacity for mitochondrial energy production, such that capacity exceeds stress-induced energy demand, could permit rapid and effective adaptation to stressors and promote timely recovery from stressors. This adaptation might prove particularly critical in situations of prolonged and repeated stressors known to cause AL (McEwen, 1998b). The following section considers factors that influence mitochondrial content and function, including behavioral and genetic factors, in the context of social inequality. We conclude by considering conditions wherein mitochondrial energetic capacity may be exceeded, leading to mitochondrial stress, mitochondrial AL, and downstream allostatic consequences such as cellular senescence (Picard, Juster, & McEwen, 2014). Behaviors Influencing Mitochondrial Function The effects of physical activity and other protective behaviors that foster biological resilience may lie in improved mitochondrial energy production. The amount of energy cells can use and produce is in large part dictated by the quality and quantity of mitochondria they contain (Schwerzmann, Hoppeler, Kayar, & Weibel, 1989). For instance, cells with more healthy mitochondria can perform more work than those with fewer mitochondria. Equally important, high mitochondrial content enables cells to sustain function at a given workload for longer periods before fatigue onset. For instance, individuals with mitochondrial diseases (where mitochondria cannot normally produce energy for muscles to contract) can exercise only at low intensities and can maintain given workloads only for limited periods of time (Taivassalo et

al., 2003). In response to increased energy demand, cells adapt by increasing the number of energyproducing mitochondria in their cytoplasm via a process called mitochondrial biogenesis (Handschin & Spiegelman, 2008). Physical activity, which increases the energy required and consumed by the contracting muscles and the working brain, triggers mitochondrial biogenesis. Resulting rises in energy demand initiate cellular signaling cascades that consequentially increase mitochondrial content in muscles (Hood, 2009), in different brain regions (Steiner, Murphy, McClellan, Carmichael, & Davis, 2011) and systemically in other organs (Little, Safdar, Benton, & Wright, 2011). Changes in mitochondrial content and morphological shape as well as other concomitant antioxidant, quality control, and anti-inflammatory processes are thought to underlie the beneficial effects of health behaviors that transiently increase energy demand, such as exercise (Handschin & Spiegelman, 2008; Picard & Turnbull, 2013). Health behaviors that increase cellular energy capacity can promote successful adaptation. Equipped with a healthy pool of well-functioning mitochondria, cells can more easily meet the energy cost of stress and thus minimize AL. In the long run, this reserve would contribute to biological resilience. In line with this, regular exercise may prevent the deleterious effects of chronic psychosocial stresses on cellular senescence as measured with telomere length (Puterman et al., 2010). It is important to recall that lifestyle factors are integral behavioral components of AL. For instance, caloric overload and excessive dietary sugar generates free radicals and inflammation. Similarly, physical activity can also contribute significantly to the association between stress/depression and cardiovascular disease, with physically active and fitter individuals appearing to be more resilient to stressors (Hamer, 2012). Likewise, aging is associated with a decline in mitochondrial content and function (Short et al., 2005), which is partially normalized by physical activity (Lanza et al., 2008). Physical activity, which stimulates mitochondrial biogenesis and affects other cellular processes that improve mitochondrial function, is therefore a potential source of cellular resilience. Conversely, factors that reduce cells' energy-producing capacity are associated with adverse health outcomes, including cardiometabolic diseases and cognitive decline (F. B. Hu et al., 2004; Mattson, 2012). Behaviors leading to reduced mitochondrial content and function include physical inactivity and excess calorie intake (overeating) or type 2 allostatic overload. Because organs require less glucose and lipids when they are inactive, food substrates accumulate systemically. This accumulation leads to the oversupply of energetic substrates (glucose and lipids) relative to demand, also known as a state of metabolic oversupply (Picard & Turnbull, 2013). Metabolic oversupply can compromise mitochondrial health in three major ways: (1) by suppressing mitochondrial biogenesis, thereby reducing the synthesis of new mitochondria and lowering of mitochondrial content within cells of the body (Petersen, Dufour, Befroy, Garcia, & Shulman, 2004); (2) by inhibiting the removal of damaged and dysfunctional mitochondrial, which normally occurs on a continuous basis and contributes to normal mitochondrial function (Liesa & Shirihai, 2013); and (3) by causing prolonged mitochondrial fragmentation (Picard & Turnbull, 2013), which leads to the accumulation of mitochondrial DNA damage (Fukagawa et al., 1999), oxidative stress, and molecules that trigger cell death (Yu, Robotham, & Yoon, 2006). Thus, behaviors that promote metabolic oversupply and

reduce mitochondrial energy-producing capacity can lead to AL. Environmental and social factors inherent to social inequalities can significantly shape one's behavior and mitochondria, at least in part via physical activity and diet. Neighborhoods and social environments that encourage the regular practice of physical activity (e.g., access to facilities, safe walking or biking paths, social engagement) or promote sedentary lifestyles (e.g., urbanized neighborhoods, unsafe environments, social desirability, stigmatization, selfperceptions, and discrimination (Sabiston & Crocker, 2008)) may exert strong indirect influences on metabolic state and thus affect biological resilience. Altered behaviors (physical activity and diet) and primary AL mediators (cortisol, cytokines) might synergistically influence mitochondrial function. Thus, psychosocial and behavioral factors linked to social inequalities may influence AL and vulnerability/resilience to disease by contributing to mitochondrial health and the organism's adaptive capacity. In addition and as is explored next, factors intrinsic to mitochondria could influence how cells adapt to stress. Mitochondrial Genome, Haplogroups, and Telomeres Essential to mitochondrial energy production is mtDNA, a circular genome residing within mitochondria that encodes only 37 genes and is inherited from the mother only (R. W. Taylor & Turnbull, 2005). Although this genome is approximately 200,000 times smaller than the nuclear genome, 98% of the mtDNA encodes essential components of energy production. Pathogenic mutations in the mtDNA therefore have severe health consequences (DiMauro & Schon, 2003). However, other “milder” genetic variations in the form of single nucleotide polymorphisms (SNPs) exist in the mtDNA and appear to have subtle effects on the function of mitochondria (Gomez-Duran et al., 2010). mtDNA SNPs have historically segregated as groups—called haplogroups—during human evolution and migration (Wallace, 2005). To date, several studies have linked mtDNA variants to disease susceptibility (Wallace, 2013), such that specific haplogroups appear to predispose to specific illnesses, such as schizophrenia and bipolar disorder (Rollins et al., 2009; Sequeira et al., 2012), AIDS progression (Hendrickson et al., 2008), and diabetes (Crispim et al., 2006). By contrast, specific mtDNA haplogroups have also been associated with athletic performance (Niemi & Majamaa, 2005) and longevity (De Benedictis et al., 1999), indicating that subtle changes in mtDNA can confer susceptibility or resistance to age-related disease. It is informative to consider the origin of human mtDNA haplogroups. mtDNA haplogroups arising 50,000 to 200,000 years ago are thought to have permitted the migration of humans out of Africa, from warmer to colder climates based on differences in mitochondrial function (Wallace, 2005). According to this theory, individuals with a certain mtDNA haplogroup would have mitochondria that are well coupled electrochemically, which makes them ideally suited to warm climates and possibly certain types of food; whereas other individuals with certain mtDNA haplogroups would be better suited to tolerate colder climates and thus could migrate out of Africa (Wallace, 2005). As a result, race/ethnic groups with ancestry originating from geographically distinct regions (e.g., Africa, Asia, North Europe, Siberia) characterized by warmer or colder climates harbor divergent mtDNA haplogroups. In other words, in addition to the sociocultural variables typically associated with race/ethnicity (e.g., cultural

differences, diet, family structures, value sets, visible minority features that are often sources of discrimination), different “types” of mitochondria could further modulate disease susceptibilities. In particular, different mtDNA haplogroups may generate mitochondria that respond differently to environmental factors, behaviors, and psychosocial factors influencing energy metabolism (Liou et al., 2007). This alteration might contribute to shape an individual's responses to chronic stress and thus AL. For instance, a specific mtDNA haplogroup or SNP may exert little or no effect under normal circumstances but can become deleterious in the presence of chronic metabolic stress (type II allostatic overload) when mitochondria are challenged (Liou et al., 2007). Studies are needed to disentangle the biological effect of mtDNA haplogroups from those of other sociocultural factors that frequently correlate with race/ethnicity. It may be that inherited mtDNA haplogroups differentially alter disease susceptibility in conjunction with social inequalities, particularly in the context of racially and culturally diverse study samples. A related biomarker related to aging is the length of telomeres, the protective DNA-protein complexes at the end of chromosomes that shorten as cells age. Telomere shortening is therefore a marker of cellular senescence (Chan & Blackburn, 2004). Interestingly, individuals exposed to chronic psychological stress (Epel et al., 2004) or with low SES (Needham, Fernandez, Lin, Epel, & Blackburn, 2012; Needham et al., 2013) have shorter telomeres in blood cells, and A mtDNA haplogroups are associated with blood cell telomere length (Fernandez-Moreno et al., 2011). This finding substantiates the notion that cellular senescence and the aging process are characterized by telomere shortening, which is also interconnected with mitochondrial dysfunction (Sahin & Depinho, 2010). Despite its popularity as a biomarker of senescence, telomere erosion can occur early in life: A recent study demonstrated that 5-year-old children exposed to heightened domestic violence displayed significantly more telomere erosion than counterparts when adjusting for sex, SES, and BMI (Shalev et al., 2013). In summary, mitochondria, mtDNA haplogroups, and telomeres are biomarkers that should be assessed in AL research on vulnerability and resilience. However, other neurobiological factors also mediate the effects of stress on AL and psychiatrically relevant phenotypes, including hippocampal neuroplasticity involved in regulating HPA activity and stress response.

Hippocampal Neurogenesis The central organ of stress physiology and AL is the brain because it determines what is threatening and then orchestrates the appropriate physiological and behavioral responses needed to cope. The brain is also a plastic and vulnerable organ and changes in the architecture of key brain regions, such as the hippocampus, PFC, amygdala, and nucleus accumbens, alter how these brain regions respond. In determining susceptibility to negative outcomes following chronic stress, it is important to consider the neurobiological underpinnings of neuroplastic responses to stress in the brain. In this respect and consistent with the life cycle model of stress, the hippocampus is particularly important as it modulates HPA axis activity through its inhibition of the hypothalamus (Herman, Ostrander, Mueller, & Figueiredo, 2005; Pruessner et al., 2005, 2008,

2010). The hippocampus is also relatively unique among brain structures in that it undergoes substantial neurogenesis (the creation and integration of new neurons) throughout life (Knoth et al., 2010). This process has particular significance to stress (Schloesser, Manji, & Martinowich, 2009; Snyder, Soumier, Brewer, Pickel, & Cameron, 2011), affective modulation (Mahar et al., 2011; Perera et al., 2011; Santarelli et al., 2003; Snyder et al., 2011; J. W. Wang, David, Monckton, Battaglia, & Hen, 2008), and psychiatric phenotypes (Boldrini et al., 2013; Perera et al., 2011; Reif et al., 2006). It may also play an important role in the mechanism by which social inequality affects AL. Hippocampal neurogenesis results in the continual addition of new granule cell neurons to the dentate gyrus (DG) of the hippocampus (see Kempermann, 2011, for review). These neurons originate from local neural stem cells, which produce rapidly amplifying neuronal precursors. Following several weeks of development, during which these cells differentiate into neurons and extend neurites, these immature neurons functionally integrate into existing DG circuits, contributing to hippocampal plasticity. In addition to constantly shaping the organization of DG circuitry, these cells also individually show increased plasticity in their firing activity relative to mature neurons (Ge, Yang, Hsu, Ming, & Song, 2007). Neurogenesis is modulated by a wide variety of factors, including exercise (Kiuchi, Lee & Mikami, 2012; Kronenberg et al., 2006; Trejo, Carro, & Torres-Aleman, 2001), sleep (Meerlo, Mistlberger, Jacobs, Heller, & McGinty, 2009; Mirescu, Peters, Noiman, & Gould, 2006; Mueller, Mear, & Mistlberger, 2011; Mueller et al., 2008), age (Knoth et al., 2010; Kronenberg et al., 2006; Kuhn, DickinsonAnson, & Gage, 1996), diet (J. Lee, Duan, Long, Ingram, & Mattson, 2000; J. Lee, Duan, & Mattson, 2002; Lindqvist et al., 2006), neurotrophic factors (Aberg, Aberg, Hedbacker, Oscarsson, & Eriksson, 2000; Jin et al., 2002; J. Lee et al., 2002; Mahar et al., 2011; Schmidt & Duman, 2010), biological sex (Boldrini et al., 2009; Chow, Epp, Lieblich, Barha, & Galea, 2012; Perfilieva, Risedal, Nyberg, Johansson, & Eriksson, 2001), antidepressant medication (Boldrini et al., 2009, 2012; Malberg, Eisch, Nestler, & Duman, 2000; Perera et al., 2011), and (of particular relevance) stress (Dagyte et al., 2011; Elizalde et al., 2010; Jayatissa, Henningsen, West, & Wiborg, 2009; Mineur, Belzung, & Crusio, 2007; Perera et al., 2011). In animal models, chronic stress induces depressivelike behavior and decreased hippocampal volume and hippocampal neurogenesis (Brummelte & Galea, 2010; Dagyte et al., 2011; Elizalde et al., 2010; Jayatissa et al., 2009; Mineur et al., 2007; Perera et al., 2011; Schmidt & Duman, 2010; Valente et al., 2012; Veena et al., 2009; Wong & Herbert, 2004). The latter is thought to result from the stimulation of GR receptors on neurogenic cells by cortisol, leading to decreased transforming growth factor-β-SMAD2/3 and hedgehog signaling and subsequent decreases in cell proliferation, neuronal differentiation, and survival (Anacker et al., 2012; Dagyte et al., 2011; Elizalde et al., 2010; Jayatissa et al., 2009; Mineur et al., 2007). Neurotrophic factors, including insulin-like growth factor (IGF; Aberg et al., 2000), vascular endothelial growth factor (VEGF; Jin et al., 2002), and neuregulin-1 (NRG1; Mahar et al., 2011), also regulate adult hippocampal neurogenesis and affective behavior. They may also govern biological resilience or susceptibility to HPA axis dysfunction and psychopathology in the context of adverse life circumstances. Notably, brain-derived neurotrophic factor (BDNF) is thought to mediate the interaction between neurogenesis and stress (Suri & Vaidya, 2012).

Indeed, hippocampal (including DG) BDNF levels: decrease following chronic stress (Shi, Shao, Yuan, Pan, & Li, 2010; Smith, Makino, Kim, & Kvetnansky, 1995a; Smith, Makino, Kvetnansky, & Post, 1995b); increase following antidepressant treatment (Altar, Whitehead, Chen, Wortwein, & Madsen, 2003; Czubak et al., 2009; Gersner, Toth, Isserles, & Zangen, 2010; Hanson, Owens, & Nemeroff, 2011; Musazzi et al., 2009; Nibuya, Morinobu, & Duman, 1995), and indirect modulate hippocampal neurogenesis and affect. Specifically, higher BDNF levels are associated with increases in neurogenesis and antidepressant effects (Scharfman et al., 2005; Schmidt & Duman, 2010; Shirayama, Chen, Nakagawa, Russell, & Duman, 2002), whereas lower levels have been associated with decreased neurogenesis, depressivelike behavior, and behavioral insensitivity to antidepressants (Adachi, Barrot, Autry, Theobald, & Monteggia, 2008; Rossi et al., 2006; Sairanen, Lucas, Ernfors, Castren, & Castren, 2005; Taliaz, Stall, Dar, & Zangen, 2010). In addition, expression of Tropomysoin receptor kinase b or TrkB, the gene encoding the BDNF receptor, is required in hippocampal neuronal precursors (but not mature granule cells) for neurogenic and behavioral responses to antidepressants (Y. Li et al., 2008). Finally, genetic alterations of BDNF or TrkB in humans have been associated with depression and altered antidepressant response as well as suicidal behavior and other psychopathology (Ernst et al., 2011, 2012; Licinio, Dong, & Wong, 2009; Perlis, Ruderfer, Hamilton, & Ernst, 2012). Thus, reduction of hippocampal BDNF levels and signaling represent an alternative putative mechanism by which increased stress reduces hippocampal neurogenesis. Underlining the complex nonlinear network of interconnected allostatic mechanisms, an emerging body of research has revealed that mitochondria can function as neuromodulators (Picard & McEwen, 2014). Indeed, the very process of synaptic transmission that constitutes the foundation of current theories for brain function and cognition is modulated by the presence of mitochondria in presynaptic terminals (Ivannikov, Sugimori, & Llinas, 2013; T. Sun, Qiao, Pan, Chen, & Sheng, 2013; M. Vos, Lauwers, & Verstreken, 2010). A recent study further found that the morphology of PFC presynaptic mitochondria—which is contingent on their function and oxidative stress—correlated with synapse morphology and memory performance in living monkeys (Hara et al., 2014). Likewise, the genetic mixing of different types of mtDNA in mice resulted in severely impaired spatial learning and memory retention capacity, indicating that subtle bioenergetic changes have significant functional consequences on brain function and cognition (Sharpley et al., 2012). It is also noteworthy that several widely used drugs to treat affective spectrum disorder affect mitochondrial function (Gardner & Boles, 2011), and newgeneration antidepressive agents also have preferential action on mitochondrial energy metabolism (Nasca et al., 2013). Mitochondrial remodeling of brain function is an emerging additional mechanism of great interest whereby environmental and biological factors can impact neuroplasticity. Septotemporal Axis and Stress Response As with the life cycle model of stress described earlier, the hippocampus has been extensively implicated in stress reactivity and HPA axis function (Herman et al., 2005; Pruessner et al., 2005, 2008; Schloesser et al., 2009), and hippocampal neurogenesis in particular has also

been implicated in these functions (Schloesser et al., 2009; Snyder et al., 2011). The involvement of hippocampal neurogenesis in HPA axis function and stress response is likely related to its anatomy and functional connectivity. Specifically, the posterior hippocampus in humans (analogous to the dorsal hippocampus in rodents) is associated with learning and memory, and projects to the lateral septum (O'Keefe & Nadel, 1978), whereas the anterior hippocampus (ventral in rodents) is more involved with affective functions, and connects to the PFC, amygdala, and, most relevantly, the paraventricular nucleus of the hypothalamus via the bed nucleus of the stria terminalis (Bannerman et al., 2004; Fanselow & Dong, 2010). Notably, long-term plasticity in the ventral hippocampus in animals is selectively affected by stress and glucocorticoids (Maggio & Segal, 2007), and lesions to areas of this subregion specifically projecting to the hypothalamus induce changes in stress regulation (Nettles, Pesold, & Goldman, 2000). Hippocampal neurogenesis has been extensively implicated in stress modulation. More specifically, chronic stress has been demonstrated to decrease neurogenesis in the ventral hippocampus (Elizalde et al., 2010; Jayatissa, Bisgaard, Tingstrom, Papp, & Wiborg, 2006; Jayatissa, Bisgaard, West, & Wiborg, 2008; Tanti, Rainer, Minier, Surget, & Belzung, 2012), whereas increases in neurogenesis restricted to the same subregion have been associated with antidepressantlike effects as well as normalization of hippocampal GR expression following stress (Mahar et al., 2011; Paizanis et al., 2010). Conversely, the ablation of neurogenesis disrupts HPA response to stress (Schloesser et al., 2009; Snyder et al., 2011) and prevents antidepressants from normalizing hippocampal HPA axis regulation in response to chronic stress (Surget et al., 2011). Taken together, these studies support the role of ventral (anterior in humans) hippocampal neurogenesis in regulating HPA axis activity via its connections to the paraventricular nucleus of the hypothalamus. This finding suggests that chronic stress-induced decreases in neurogenesis decrease hippocampal suppression of the HPA axis, leading to increased GC secretion in response to stress, further suppressing neurogenesis, causing HPA dysregulation and ancillary deleterious effects. By extension, an initial deficit in neurogenesis could predispose individuals to HPA axis dysregulation following chronic stress, and proneurogenic factors could act as prophylaxis against neurogenesis-related HPA axis dysregulation following stress. If individuals experiencing social inequality are indeed vulnerable to HPA axis dysregulation, as discussed, they in particular might benefit from proneurogenic interventions, such as antidepressant medication or exercise, in order to prevent or reverse neurogenesis deficit–related HPA axis dysregulation. Thus, prestress levels of hippocampal neurogenesis may serve as a vulnerability/resiliency factor to the stress-related effects of social inequalities and other diverse stressors. Social Inequalities and Neurogenesis To the extent that social inequality affects HPA axis activity, the effects of social inequality on AL are likely exacerbated through the suppression of hippocampal neurogenesis and resulting HPA axis dysregulation. As such, it is critical to determine the extent to which social inequality increases glucocorticoid secretion and HPA axis dysregulation. Animal models of social stress

have revealed that social defeat decreases hippocampal neurogenesis (see Schoenfeld & Gould, 2012 for review) and induces submissive behavior (Vivian & Miczek, 1999). In addition, environmental enrichment facilitates recovery from social stress–induced submissive behavior through a neurogenesis-dependent mechanism (Schloesser, Lehmann, Martinowich, Manji, & Herkenham, 2010). Although it remains to be demonstrated in human studies that social inequality itself is directly associated with decreased hippocampal neurogenesis, human social inequality and societal disadvantage have been shown to provoke abnormal physiological stress responses and HPA axis dysregulation. These social adverities have included, for example, low childhood SES (Lupien et al., 2001), childhood abuse (Cicchetti & Rogosch, 2001; Cicchetti et al., 2010), low maternal care (Pruessner, Champagne, Meaney, & Dagher, 2004), psychiatric illness (Staufenbiel, Penninx, Spijker, Elzinga, & van Rossum, 2012), low self-esteem (Pruessner, Hellhammer, & Kirschbaum, 1999; Pruessner et al., 2005), caregiver burden (Davis et al., 2004; de Vugt et al., 2005), minority sexual orientation (Huebner & Davis, 2005; Juster et al., 2013b), and perceived racial discrimination (Fuller-Rowell, Doan, & Eccles, 2012; Zeiders, Doane, & Roosa, 2012). One mechanism by which social inequality may affect AL is by decreasing hippocampal neurogenesis (particularly in the anterior hippocampus). This would ultimately lead to decreased inhibition of the HPA axis (specifically via the paraventricular nucleus of the hypothalamus), resulting in increased glucocorticoid release and further HPA axis dysregulation through neurogenic and nonneurogenic mechanisms. In conclusion, human hippocampal neurogenesis in the specific context of social inequalities has yet to be examined. This is due in part to the difficulty of assessing neurogenesis in vivo as well as a paucity of postmortem samples from individuals characterized by their stressful exposure to chronic social inequality. However, as techniques for measuring neurogenesis advance and as large-scale longitudinal cohorts potentially lead to the collection of more thoroughly characterized postmortem brain samples, future studies should be well positioned to confirm or revise the hypothesis that the effects of social inequality on subsequent HPA axis dysregulation are exacerbated or mediated by decreases in hippocampal neurogenesis. Further research into hippocampal neurogenesis may also contribute to our understanding of the effects of glucocorticoid neurotoxicity on hippocampal volume and cognitive functioning in elders (Lupien et al., 1998) as well as the factors driving morphological hippocampal changes throughout life (Lupien et al., 2007; Lupien et al., 2009). Along these lines, our next and final section of this part of the chapter explores the concept of cognitive reserve as a form of cognitive resilience in old age and is of importance for future AL research.

Cognitive Reserve Aging is often associated with cognitive impairment and worsening health that is consistently associated with increasing AL. Yet a considerable degree of variability in cognitive performance as well as health status exists among older adults (Mattson, Chan, & Duan, 2002; Rowe & Kahn, 1997; Wilson et al., 2002a). Epidemiological studies have consistently shown that life experiences are associated with risk for cognitive impairments (Fratiglioni, PaillardBorg, & Winblad, 2004; Scarmeas, Levy, Tang, Manly, & Stern, 2001; Stern, 2012). To

address this probabalistic process the notion of cognitive reserve helps account for interindividual differences in the risk for dementia in the presence of brain pathology and the various factors that may allow one's brain to respond appropriately to pathology and postpone the onset of symptoms (Stern, 2002; Whalley, Deary, Appleton, & Starr, 2004). It is believed that individuals with more cognitive reserve are able to tolerate more pathology prior to the appearance of symptoms, representing an important form of cognitive resilience in the face of aging and AL. This adaptation may occur through the use of neuronal compensatory mechanisms and the use of existing cognitive processes established by the individual (Stern, 2002), which we explore in this section before concluding with clinical and social implications. Various factors throughout life span development contribute to enhanced cognitive reserve that buffers the adverse biological effects of senescence. First, higher educational attainment consistently plays a protective role in epidemiological studies, whereas low educational attainment increases the risk for dementia and Alzheimer's disease (Karp et al., 2004; Launer et al., 1999; Ngandu et al., 2007; Ott et al., 1995; Qiu, Backman, Winblad, Aguero-Torres, & Fratiglioni, 2001). Similarly, benefits are observed among healthy older adults, such that those with higher educational attainment show a slower rate of cognitive decline compared to those with fewer years of formal education (Chodosh, Reuben, Albert, & Seeman, 2002; Cullum et al., 2000). As described earlier, this finding is consistent with a large body of research linking lower SES to higher AL among different social groups throughout the life span. An important factor that contributes to cognitive reserve is occupational attainment and skill sets exercised during working adulthood. Specifically, job complexity is an extremely important factor in light of the extensive amount of time people engage in work and workrelated activities throughout adulthood. Some studies show that jobs consisting of manual work increase the risk for dementia in comparison to nonmanual jobs (Bonaiuto et al., 1995; Mortel, Meyer, Herod, & Thornby, 1995). The nature of job demands may also protect against dementia and Alzheimer's disease; those requiring the use of extensive vocational training, academic skills, and complex work are protective due to their high mental demands (Andel et al., 2005; Kroger et al., 2008; Potter, Helms, Burke, Steffens, & Plassman, 2007), while results are mixed when it comes to job complexity (Andel et al., 2005; Kroger et al., 2008; Potter et al., 2007). Moreover, the risk for cognitive decline is higher among individuals with both low educational and occupational attainment (Stern et al., 1994). In line with this finding, a recent study revealed that older women with low psychological demands at work experience elevated levels of AL (Juster et al., 2013a). Taken together, demanding jobs may function as a form of mental exercise, consistent with the use-it-or-lose-it principle (Katzman, 1995). Although educational and occupational attainment are factors that occur earlier during working adulthood, experiences later in life can also play a marked role in decreasing the risk for dementia (Stern, 2009). For example, participating in leisure activities—especially those that are cognitively stimulating—can promote cognitive capacities and reduce the risk for dementia (Crowe, Andel, Pedersen, Johansson, & Gatz, 2003; Scarmeas et al., 2001; Wilson et al., 2002b). Benefits are also observed among patients with Alzheimer's disease (Friedland et al., 2001; Kondo, Niino, & Shido, 1994). A systematic review demonstrates that the protective

effects of leisure activities are observed even after controlling for factors such as age, health status, education, and occupation (Valenzuela & Sachdev, 2006). Importantly, engaging in leisure activities postpones cognitive impairment and slows the rate of decline among individuals who eventually develop dementia (Crowe et al., 2003; Scarmeas, Albert, Manly, & Stern, 2006; Stern, Albert, Tang, & Tsai, 1999; Wilson et al., 2002b). Despite the strong evidence for the benefits of higher cognitive reserve, it is noteworthy that patients with Alzheimer's disease who have higher cognitive reserve show accelerated progression of symptoms and more rapid cognitive decline (Scarmeas et al., 2001, 2006; Stern et al., 1999). This finding suggests that cognitive reserve might play an important role in “masking” existing pathology while postponing the appearance of clinical symptoms until the underlying pathology becomes more severe (Stern, 2006). Future research could investigate this paradox further by assessing how cognitive reserve manifests among diverse social strata and whether this mechanism corresponds with specific biological signatures of AL and thresholds of severity linked to neuropathology and clinical symptoms. Neuroimaging Cognitive Reserve Imaging studies using various techniques support the protective role of cognitive reserve. The first studies investigated cerebral blood flow, which decreases among patients with Alzheimer's disease (Hirao et al., 2005; Kogure et al., 2000). Findings showed that patients with higher cognitive reserve (based on educational attainment) showed dampened overall cerebral blood flow compared to patients matched for clinical severity (Stern, Alexander, Prohovnik, & Mayeux, 1992). Similar findings have been shown as a function of occupational attainment and leisure activities (Alexander et al., 1997; Perneczky et al., 2006; Scarmeas et al., 2003; Stern et al., 1995). Collectively, these studies indicate that those with higher cognitive reserve are able to resist senescence while showing similar clinical symptoms (Alexander et al., 1997; Bennett et al., 2003; Perneczky et al., 2006; Stern, 2009). Education, occupational attainment, and leisure activities may also interact with each other to enhance the protective role they each play individually. Functional magnetic resonance imaging (fMRI) studies investigating neural compensatory mechanisms have found two separate networks used when performing a working memory task. The more efficient structure, the left precentral gyrus, is used by both young and older adults, whereas the less optimal structure, in the left precentral region of the supplementary motor area, is used only by older adults and is predictive of poorer performance (Steffener, Brickman, Rakitin, Gazes, & Stern, 2009). This finding is hypothesized to be due to the fact that the first network no longer functions as effectively during aging, and this fact is partially compensated for through the increased use of the second network; however, it does not necessarily result in comparable memory performance (Stern, 2012). It also appears that individuals with higher cognitive reserve can withstand more atrophy in the first network and maintain cognitive performance, avoiding or delaying the need to resort to the second network (Steffener, Reuben, Rakitin & Stern, 2011). Among users of the second network, those with higher cognitive reserve show better cognitive performance than those with low cognitive reserve (Steffener et al., 2011). Whereas higher cognitive reserve is associated with less

activation when performing a memory task in healthy older adults, more activation was observed among individuals with mild cognitive impairment and Alzheimer's disease (SolePadulles et al., 2009). In sum, more effective neural mechanisms are present among those with higher cognitive reserve (Stern, 2012). Structural MRI studies also tend to support that notion of cognitive reserve. Importantly, atrophy of the hippocampus is a marker for dementia and Alzheimer's disease (Barnes et al., 2009). In one longitudinal study, more frequent lifelong engagement in complex activities was associated with a lower magnitude of hippocampal atrophy measured 3 years later (Valenzuela, Sachdev, Wen, Chen, & Brodaty, 2008). In one longitudinal study, more frequent lifelong engagement in complex activities in working adulthood was associated with a lower magnitude of hippocampal atrophy over time (Valenzuela et al., 2008). Such activities were related to education, occupation, reading and writing, creative arts, socializing, and daily habits during young adulthood (13–30 years of age), middle age (30–65 years), and older adulthood (65 years and above). Similarly, hippocampal-dependent neuroplasticity is found among London, UK, taxi drivers who undergo intensive training in order to navigate through complex routes (Maguire et al., 2000). In a landmark study, their posterior hippocampi were found to be larger than those of age-matched controls. These studies provide parallel evidence for the beneficial role of experience-dependent hippocampal plasticity that may potentiate cognitive reserve. Cognitive reserve has also been examined in relation to beta-amyloid, a protein that appears to accumulate in Alzheimer's disease (Ghiso & Frangione, 2002). Patients with higher educational attainment perform better than expected on cognitive tasks than would be otherwise expected considering their elevated beta-amyloid levels (Roe et al., 2008), highlighting a protective pathway against dementia-related pathophysiology (Kemppainen et al., 2008). This is an excellent example of biological and cognitive resilience as promoted by higher educational attainment that could be explored further in future AL research on geriatric neurobiology. Further identifying protective factors against AL and disease will illuminate our understanding of trajectories toward resilience to age-related diseases. Another important recurring protective factor in this chapter that may prevent or delay dementia onset is physical exercise. The performance of aerobic activities among healthy older adults can prevent cognitive impairment (Kramer & Erickson, 2007) and may therefore represent a factor that promotes cognitive reserve and biological resilience. Whereas epidemiological studies demonstrate that inactivity increases the risk for cognitive impairment and Alzheimer's disease (Andel et al., 2008; Larson et al., 2006), evidence from clinical randomized trials shows that aerobic activities can reverse cognitive decline prior to the development of dementia (Colcombe et al., 2006; Kramer et al., 1999). Consistently, metaanalyses and reviews on exercise find that it improves cognitive performance among individuals with mild cognitive impairment and Alzheimer's disease. Specific cognitive benefits have been observed for memory performance, attention, executive functioning, and processing speed (Heyn, Abreu, & Ottenbacher, 2004; P. J. Smith et al., 2010). Neuroimaging studies likewise reveal that aerobic fitness is associated with larger

hippocampal volume and better memory performance among older adults (Erickson et al., 2009). In one study, exercise training increased hippocampal volume over the course of a year, whereas in the control group, a decrease in hippocampal volume was observed (Erickson et al., 2011). Interestingly, the augmentation in hippocampal volume was associated with higher levels of BDNF (a marker that tends to be reduced among Alzheimer's patients; Peng, Wuu, Mufson, & Fahnestock, 2005). Aerobic exercise also increases gray and white matter volume in the PFC and improves executive functioning (Colcombe et al., 2004, 2006; Rosano et al., 2010) as well as whole-brain volume among Alzheimer's patients (Burns et al., 2008). These studies demonstrate that aerobic exercise has important benefits for cognitive aging. In the broader context of resilience, it would be very interesting for future research to assess whether these associations are observed in older adults exposed to adverse life conditions. Another important factor recurring throughout this chapter is social support, which consists of a quantitative component (the number of individuals in one's social network) and a qualitative component (the quality of relationships, frequency of socialization, and the perception of available social and emotional support). Both dimensions of one's social network protect against cognitive impairment in senescence (Seeman, Lusignolo, Albert, & Berkman, 2001a). Large community-based studies show that having few social contacts and, in particular, being unsatisfied with the quality of contacts is associated with increased risk for dementia (Fratiglioni, Wang, Ericsson, Maytan, & Winblad, 2000), whereas more frequent participation in social and mental activities decreases risk for dementia (H. X.Wang, Karp, Winblad, & Fratiglioni, 2002). Additional support for this association has been provided by various longitudinal studies (Bassuk, Glass, & Berkman, 1999; Fratiglioni et al., 2004; Seeman et al., 2001a). In conclusion, higher cognitive reserve decreases the risk for dementia (by up to 50%; Valenzuela & Sachdev, 2006). Importantly, while formal education and occupational attainment may be relatively fixed by older adulthood, findings demonstrate that engagement in stimulating cognitive and social activities in older adulthood can independently play protective roles (Valenzuela & Sachdev, 2006). Collectively, evidence suggests that engagement in social, mental, and physical activities, in addition to having adequate social support, can protect against dementia in late life, and this finding provides an important avenue for prevention efforts (Fratiglioni et al., 2004; Stern, 2012). The concept of cognitive reserve represents an important construct in our discussion of resilience that we believe future AL studies could help progress by further assessing specific biological processes involved.

Conclusions Clinical Implications The AL model has the potential to become a powerful tool in clinical contexts. Perhaps the most promising approach would involve tracking changes over time in AL to examine the efficacy of clinical interventions using transdisciplinary perspectives informed by triangulated combinations of biometrics, psychometrics, and sociometrics. As future studies further our

understanding of which configurations of risk factors and protective factors potentiate pathways toward resilience, it will be imperative to design prospective studies tailored to reduce AL. To date, however, no treatment options aimed at lowering AL levels exist. Nevertheless, many therapeutic options could be implemented based on the antecedents associated with AL and its clinical repercussions (see Figure 8.1). The vast array of identified risk factors and protective factors associated with vulnerability toward or resilience against AL provide scientist-practitioners with a wealth of information to help guide targeted interventions in the future. Notwithstanding, interventions attempting to reverse developmental trajectories from vulnerability toward resilient pathways may be much more complicated than simply switching risk factors into protective factors at micro- to macro-levels (e.g., hypercortisolism to hypocortisolism, physical inactivity to activity, low to high SES, etc.). Although it cannot be denied that an accumulation of protective factors promotes health and well-being, resilience as defined from a developmental psychopathology perspective implicates some history of risk factors related to adversity, stress, and/or trauma that have been modulated to allow individuals to thrive and survive, even if specific adaptations can come at a cost in other domains (e.g., academic performance and AL in John Henryism). This reality highlights the importance of trandisciplinary approaches to interventions orchestrated by teams of professionals (e.g., nurses, psychologists, psychiatrists, social workers, nutritionists, educators, etc.). A key question outlined by McEwen (2007, 2012) is whether approaches to reducing AL levels are best tailored in relation to top-down modifications versus conventional biomedical remediation. At an individual level, brain-centered programs that encourage improved sleep quality and quantity, social support, sense of purpose, self-esteem, healthy diet, substance avoidance, and physical activity would be undoubtedly beneficial. Interventions at social levels might include policies that create incentives for beneficial or best practices in the workplace, cleaner and safer neighborhoods, and encouraging advanced education. Specifically for the elderly, well-being therapy emphasizing autonomy, purpose in life, personal growth, positive relations with others, environmental mastery, social activities, and self-acceptance could be coupled with more traditional cognitive and physical interventions. Assessing changes in AL provides a powerful, multisystemic means to assess salubrious processes in a comprehensive manner amenable to clinical investigation (McEwen, 2012). To the best of our knowledge, no studies have assessed intervention effects on AL. Notably in an analysis of unaccountable changes in AL levels in the MacArthur cohort over the span of 2 to 5 years, deceases in AL were significantly related to reductions in all-cause mortality regardless of one's sex (Karlamangla, Singer, & Seeman, 2006). An important consideration is whether any decrease in AL is random, or whether could there be distinct neuroendocrine, immune, metabolic, or cardiovascular system(s) driving downward and/or upward shifts. It will be important for future studies to determine what combinations of health-promoting and damaging factors correspond to AL and systemic clustering at different ages and among different social groups that cope with stress in diverse ways. As described in this chapter, the direction of these interacting risk and protective factors is complex and at times contradictory.

For instance, one report found that African Americans—but not Caucasians—who selfreported increased environmental stressors and engaged in more unhealthy behaviors (e.g., overeating) had lower levels of depression but an overall greater risk of chronic physical conditions (Jackson, Knight, & Rafferty, 2010). Thus, it is possible that unhealthy behaviors have different modulatory effects depending on sociocultural factors. Identifying salubrious pathways could help guide health care approaches that are recognizing of the fact that one glove does not fit all. Throughout life span development, behavioral modification is complex and multifaceted but ultimately essential for reducing AL. In the context of the obesity epidemic, for example, interacting factors such as poor sleep hygiene, sedentary behaviors, and pollution are ubiquitous in modern societies and can compromise the efficacy of dieting among obese individuals (Tremblay & Chaput, 2012). Such unhealthy behaviors can also exacerbate interacting genetic and behavioral vulnerabilities. For instance, a novel study of Puerto Rican adults living in the United States showed that genetic polymorphisms related to lipid metabolism were only related to AL in interaction with unhealthy dietary intake (Mattei, Bhupathiraju, & Tucker, 2013). This finding has important implications for advancing our understanding of acculturation effects among different groups that can be compounded by intersecting social pressures. By the same logic, identifying and informing individuals with genetic vulnerabilities that specific diets are especially dangerous based on their constitution and circumstances is an important preventive approach that must be considered in the evolution of person-centered health care and consultation. Given cultural differences in dietary intake, and given the links among genetics, metabolism, and diet to name a few, future studies could assess associations among stress, diets, and AL in various social groups and between the sexes over life span development. Another important health behavior is sleep-related behaviors that regulate allostatic mechanisms (e.g., circadian clocks) and AL levels (Karatsoreos & McEwen, 2011; McEwen, 2006). Following sleep restriction, elevations in AL biomarkers occur (decreased glucose tolerance, increased evening cortisol, increased sympathetic activity); however, these allostatic processes are particularly pronounced for low-SES individuals who consequently experience increased risk of sleep problems, obesity, diabetes, and hypertension (Van Cauter & Spiegel, 1999). More broadly, researchers studying stress-related phenomena should account for sleep, given its important bidirectional association to psychiatric symptomatologies. Measuring AL in such investigations would provide important information for stress and sleep researchers alike who are studying overlapping mechanisms. Advances in neuroscience are also providing compelling insights into complex pathways of biological resilience wherein brain plasticity is modulated by behavior and environment (McEwen & Gianaros, 2011). For example, French stroke patients treated early with antidepressant medication in conjunction with physiotherapy experience better recovery of motor deficits over time (Chollet et al., 2011). In the spirit of “stress inoculation,” short periods of food restriction in rats predictably increases stress hormone levels, but in a positive manner that can help cure amblyopia (lazy eye) otherwise considered irreversible (Chollet et al., 2011). Whereas the mechanisms involved require further elucidation, the concept of

hormesis, that is, evolutionarily conserved biological patterns wherein mild stressors can buffer against noxious effects of subsequent stressors (Pijl, 2012), provides some insights. Studying how hormesis might function vis-à-vis mitochondrial sustainability, neurogenesis proliferation, or cognitive reserve and other biopsychosocial processes that ultimately foster resilience is a fascinating frontier for future AL research. As AL represents multisystemic, subclinical physiological dysregulations often predating the emergence of clinical outcomes, treatment options targeting psychological, behavioral, cognitive, and social domains discussed in this chapter could be alternatives or complements to biomedical approaches (Juster & Lupien, 2012b). These clinical approaches could perhaps even improve compliance and facilitate proactive prevention instead of reactive prescription. Indeed, we strongly believe that implementing person-centered approaches to diminishing AL collectively represents viable alternatives or complementary approaches to pharmacological strategies. For instance, the effects of chronic stress can be improved via pharmaceutical agents such as sleep medications, anxiolytics, antidepressants, and beta-blockers as well as by drugs that reduce oxidative stress and inflammation such as statins, insulin resistance treatments, and painkillers (McEwen, 2007). Yet there are counterarguments against pharmaceutical remediation, especially as polypharmacy relates to the widespread systemic dysregulations inherent in AL. In Sterling and Eyer's (1988) original formulation of the allostasis concept, they argued that medical practices based on homeostatic models were in danger of potentially damaging side effects and polypharmacy, as treatment problems can arise when correcting one parameter causes other systems to become dysregulated (Sterling, 2004). Allostatic regulatory systems interact in a hierarchical, dynamic, and nonlinear manner in response to stressors that prompt compensatory behaviors. From this perspective, lifestyle choices can be considered as “higher-order” allostatic mechanisms activated in reaction to external and internal cues. As such, maladaptive behaviors have the propensity of increasing AL levels even further. Taken together, unhealthy lifestyle behaviors may often represent behavioral compensatory mechanisms used to counteract negative affective states related to inherent neurobiological predispositions. From this perspective, these coping behaviors are allostatic mechanisms because they compensate for dysfunction, but ultimately at the expense of long-term health. With this in mind, therapeutic interventions targeting either specific neural pathways or the substitution of damaging behaviors for less or nondamaging ones (e.g., physical exercise, dietary restriction) would be beneficial. In addition, several psychotropic agents seem to interact with illness-related predispositions to synergistically worsen some unhealthy lifestyle behaviors and ultimately induce a type of “pharmacological” AL. As first described by our group (Juster et al., 2011), pharmacologicalcal AL (PAL) refers to the adverse effects medications can exert that then inadvertently prompt individuals to remediate the biological system(s) affected by unhealthy means . For example, smoking is often used to dampen adverse side effects of antipsychotic medications among those with severe mental illnesses, such as schizophrenia or bipolar disorder, but this compensatory behavior further amplifies AL levels (Bizik et al., 2013). Such behavioral side effects represent maladaptive allostatic mechanisms whereby individuals strive to reregulate medicinally altered neurotransmitter functions linked, for instance, to

cognition and motivation. We propose that using AL indices could provide insights into PAL in the context of psychiatric remediation (Bizik et al., 2013). To quote Jackson and colleagues (2010): “[S]uccessful interventions to reduce the use of unhealthy coping behaviors over the life course among populations living under chronically stressful conditions depend upon the recognition that such behaviors may have adaptive, neurological effects that alleviate negative psychological and physiological states.” (p. 7). The PAL concept goes beyond the known side effects of psychopharmacology; it endeavors to understand how interventions meant to treat primary diagnoses inadvertently prompt physiological dysregulations and comorbidities via patients' behavioral compensations. We propose that applying the AL model to the interpretation of the results from standard laboratory reports can provide medical professionals with a sensitive tool to monitor patients' global response to treatment (Bizik et al., 2013; Juster & Lupien, 2012b). The PAL concept also constitutes an advantage compared to focusing solely on numerous known side effects because it provides a global measure that subsumes the direct effects of the primary pathology, treatment, the effects of maladaptive compensatory behaviors, as well as other as-yet unidentified contributors to AL. Given that many of the biomarkers collected in routine blood work are the same as those used in constructing AL indices, applying simple calculations could be easily integrated into a typical clinical follow-up (Juster & Lupien, 2012b). AL indices can also serve as a starting point for a discussion between the patient and the health practitioner regarding adaptive behaviors and prevention of the development of negative sequelae related to the illness and its treatment (Bizik et al., 2013). Moreover, from an empirical standpoint, it is not yet known what constitutes critical levels of AL and how such thresholds could be used to predict specific stress-related conditions. For example, does an AL score greater than 5 represent a critical threshold for developing cardiovascular disease, and which biomarkers are more important? Because the AL index is an objective reflection of biological functioning that is intricately interconnected with genetic, neurological, developmental, behavioral, cognitive, and social factors, AL captures complex interactions that cannot be easily deduced until clinical research is conducted. Ideally, the richness of data collected in clinical practice by various professionals could be compiled and explored in relation to AL levels. All individuals have the capacity for considerable resilience in the face of adversity via beneficial processes and the appropriation of protective factors. In identifying salubrious pathways, greater appreciation must be exercised in our search for the biological mechanisms that drive resilience and how these can be harnessed to improve health (McEwen, 2003). The stress of inequality may manifest at a cellular or subcellular level, and understanding this relationship may drive further efforts in preventing and treating the pernicious effects of inequality on AL and its clinically relevant phenotypes. As explored next at a societal level, reducing social inequalities should also minimize conditions that elevate AL, thus preventing disease onset or progression.

Social Policy Implications

The AL model could be used at a macro-level to inform public health and the implementation of social policies. Daniels, Kennedy, and Kawachi (2000) have argued that reducing socioeconomic gradients and providing equal opportunities to all members of society will produce dramatic health benefits. Using AL algorithms to assess the efficacy of an intervention or policy could be a powerful way to substantiate this proposal. AL increases the risk of developing morbidities that currently burden population health worldwide to a disproportionate degree. The array of mediators and moderators constituting AL represents a physiological mechanism at the heart of socially determined health inequalities. We propose that knowledge derived from the AL model can provide policymakers with tools to help prioritize and substantiate tentative public health interventions aimed at reducing social inequalities in population health. By quantifying changes in AL over time, policymakers can also determine the efficacy of such policies. In so doing, public health officials may curb growing inequalities and monitor progress in programs that promote health. Ultimately, the biological sciences need to provide policymakers with the much-sought-after causal links for these effects and with the tools necessary to effectively evaluate health and well-being (McEwen, 2001). Our overview on social inequalities influencing AL provides a biological bridge linking social injustice to health inequalities. Although social contexts are generally measured using aggregates of individuals, diseases occur at the level of the individual, making it very difficult to assess the impact of social policy implementation (Thisted, 2003). Advances in biological sciences should therefore be applied to tease apart the factors that contribute to the paradox that while we generally live longer, health and wealth are shared so unequally. Policies often fail to attend adequately to determinants of health upstream from the provision of medical interventions (Daniels, Kennedy, & Kawachi, 2000) like those discussed in our previous section on clinical implications. Furthermore, there is growing recognition within numerous disciplines that significant risk factors associated with developing numerous pathologies are socially determined (Marmot, 2005). Likewise, Daniels, Kennedy, and Kawachi (2000) proposed that suboptimal aggregate health measures result from unnecessary differentials in the distribution of resources among social classes. The current means by which societies allocate resources within their populations is therefore a key contributor to socioeconomic gradients in disease (Daniels et al., 2000). As social inequalities are bad for our health and are socially constructed, differentials in health measures would undoubtedly lessen through a more equal reorganization of society. Knowledge generated by the AL model at theoretical and empirical levels expands epidemiological evidence of population health inequalities. Complementarily, social justice focuses on the premise of equality of opportunity. Social and political institutions should mitigate the effects of social inequalities that prevent less fortunate members of society from obtaining equal health opportunities and chances to flourish in life. Importantly, there are potential critical periods to identify and prevent individuals from developing elevated AL. Indeed, findings from the large-scale NHANES reveal that AL steadily increases with age up through the 20s to 60s and then plateaus throughout the 60s to 90s, during the period of greatest mortality risk (Crimmins et al., 2003). From life course

perspectives, these trajectories represent decades-long windows of opportunity to intervene before individuals succumb to disease and death. Could there be windows of opportunity for best periods in life to develop resilience in specific life domains that could minimize AL? Despite sensitive periods that occur early on in life during critical windows of development, positive health can be achieved at any age (Cicchetti, 2013). The best way to assess the real impact of interventions on AL would be to measure their potential for changes at different ages, different sexes, different SES, and different cultures internationally. It is also important to recognize the potential to improve health and well-being at every life stage. For example, a fMRI study of African American older women of lower SES who were at risk of cognitive deterioration showed improvements in left PFC and anterior cingulate cortex functioning matched by behavioral improvements in executive functioning 6 months after volunteering in the social service Experience Corp. program compared to waitlist controls (Carlson et al., 2009). This finding has important implications for the implementation of interventions that can target social inequalities and age-related cognitive deterioration even in late life. To summarize the implications for the field of developmental psychopathology, the AL model has proven powerful in explaining the pervasive and long-lasting consequences of social inequalities on physical and mental health. In many ways, the AL model adopts numerous principles guiding developmental psychopathology. For example, it focuses on bidirectional transactions occurring between the individual and the environment, and it proposes progressive rather than discrete transitions between normal/adaptive responses (allostasis) and pathological states (AL and allostatic overload). Moreover, it emphasizes the cumulative and synergic impacts of changes occurring at multiple levels of analysis, including verticaland horizontal-level interactions. Furthermore, the AL model follows transdisciplinary perspectives (Juster et al., 2011) and process-oriented approaches that provide explanations for underlying differences between individuals exposed to similar life circumstances (multifinality). All of these constructs are considered conceptual landmarks in developmental psychopathology (Cicchetti, 2013). In conclusion, the AL model originated with the study of vulnerabilities to stress-related disease and has recently evolved to include approaches toward greater understanding of resilience. This paradigm shift is important for three key reasons. First, AL studies of resilience are consistent with a central tenet in developmental psychopathology over and above those just mentioned; that is, that probabilistic pathways toward vulnerability or resilience fall along a continuum over life span development. Second, a focus on resilience to AL allows researchers to further identity protective factors integral to biological integrity despite preexisting risk factors associated with life histories of stress, strain, and stigma. And third and most broadly, scientific investigations that can reveal success stories among socially marginalized groups represent an important platform from which to change social perspectives. If the majority of individuals can learn of the remarkable resilience exemplified by hardy individuals exposed to social inequalities, a connectedness between social groups is fostered. This connection is otherwise more difficult to achieve with doom-and-gloom perspectives focused solely on vulnerabilities to AL since it overlooks the inspiring potential among all

people to change their destiny.

References Aberg, M. A., Aberg, N. D., Hedbacker, H., Oscarsson, J., & Eriksson, P. S. (2000). Peripheral infusion of IGF-I selectively induces neurogenesis in the adult rat hippocampus. Journal of Neuroscience, 20, 2896–2903. doi: 0270–6474/00/202896–08 Adachi, M., Barrot, M., Autry, A. E., Theobald, D., & Monteggia, L. M. (2008). Selective loss of brain-derived neurotrophic factor in the dentate gyrus attenuates antidepressant efficacy. Biological Psychiatry, 63, 642–649. doi: 10.1016/j.biopsych.2007.09.019 Adler, N. E., Epel, E. S., Castellazzo, G., & Ickovics, J. R. (2000). Relationship of subjective and objective social status with psychological and physiological functioning: preliminary data in healthy white women. Health Psychology, 19, 586–592. doi: 10.1037/0278–6133.19.6.586 Adler, N., & Snibbe, A. C. (2003). The role of psychosocial processes in explaining the gradient between socioeconomic status and health. Current Directions in Psychological Science, 12, 119–123. doi: 10.1007/s00059–013–4040–7 Alexander, G. E., Furey, M. L., Grady, C. L., Pietrini, P., Brady, D. R., Mentis, M. J., & Schapiro, M. B. (1997). Association of premorbid intellectual function with cerebral metabolism in Alzheimer's disease: Implications for the cognitive reserve hypothesis. American Journal of Psychiatry, 154, 165–172. Allsworth, J. E., Weitzen, S., & Boardman, L. A. (2005). Early age at menarche and allostatic load: Data from the Third National Health and Nutrition Examination Survey. Annals of Epidemiology, 15, 438–444. doi: 10.1016/j.annepidem.2004.12.010 Altar, C. A., Whitehead, R. E., Chen, R., Wortwein, G., & Madsen, T. M. (2003). Effects of electroconvulsive seizures and antidepressant drugs on brain-derived neurotrophic factor protein in rat brain. Biological Psychiatry, 54, 703–709. doi: 10.1017/S1461145712000053 Anacker, C., Cattaneo, A., Luoni, A., Musaelyan, K., Zunszain, P. A., Milanesi, E.,…Pariante, C. M. (2012). Glucocorticoid-related molecular signaling pathways regulating hippocampal neurogenesis. Neuropsychopharmacology, 38, 872–883. doi: 10.1038/npp.2012.253 Anda, R. F., Felitti, V. J., Bremner, J. D., Walker, J. D., Whitfield, C., Perry, B. D.,…Giles, W. H. (2006). The enduring effects of abuse and related adverse experiences in childhood. A convergence of evidence from neurobiology and epidemiology. European Archives of Psychiatry and Clinical Neurosciences, 256, 174–186. doi: 10.1007/s00406–005–0624–4 Andel, R., Crowe, M., Pedersen, N. L., Fratiglioni, L., Johansson, B., & Gatz, M. (2008). Physical exercise at midlife and risk of dementia three decades later: A population-based study of Swedish twins. Journals of Gerontology: Series A, Biological Sciences and Medical Sciences, 63, 62–66.

Andel, R., Crowe, M., Pedersen, N. L., Mortimer, J., Crimmins, E., Johansson, B., & Gatz, M. (2005). Complexity of work and risk of Alzheimer's disease: A population-based study of Swedish twins. Journals of Gerontology: Series B, Psychological Sciences and Sociological Sciences, 60, P251–P258. doi: 10.1093/geronb/60.5.P251 Arseneault, L., Milne, B. J., Taylor, A., Adams, F., Delgado, K., Caspi, A., & Moffitt, T. (2008). Being bullied as an environmentally mediated contributing factor to children's internalizing problems: A study of twins discordant for victimization. Archives of Pediatric and Adolescent Medicine, 162, 145–150. doi: 10.1001/archpediatrics.2007.53 Australian Bureau of Statistics. (2006). National Aboriginal and Torres Strait Islander Health Survey: Australia, 2004–05. Canberra, Australia: Author. Australian Bureau of Statistics. (2007). Population distribution, Aboriginal and Torres Strait Islander Australians, 2006. Canberra, Australia: Author. Australian Institute of Health and Welfare. (2008). Cardiovascular disease and its associated risk factors in Aboriginal and Torres Strait Islander peoples 2004–05. Canberra, Australia: Author. Australian Institute of Health and Welfare. (2011). Life expectancy and mortality of Aborigal and Torres Strait Islander people. Canberra, Australia: Author. Ayotte, P., Carrier, A., Ouellet, N., Boiteau, V., Abdous, B., Sidi, E. A.,…Dewailly, E. (2011). Relation between methylmercury exposure and plasma paraoxonase activity in Inuit adults from Nunavik. Environmental Health Perspectives, 119, 1077–1083. doi: 10.1289/ehp.1003296 Bagot, R. C., & Meaney, M. J. (2010). Epigenetics and the biological basis of gene × environment interactions. Journal of the American Academy of Child & Adolescent Psychiatry, 49, 752–771. doi: 10.1016/j.jaac.2010.06.001 Baker, P. T., Hanna, J. M., & Baker, T. S. (1986). The changing Samoans: Behavior and health in transition. New York, NY: Oxford University Press. Bannerman, D. M., Rawlins, J. N., McHugh, S. B., Deacon, R. M., Yee, B. K., Bast, T.,… Feldon, J. (2004). Regional dissociations within the hippocampus—memory and anxiety. Neuroscience and Biobehavarioral Reviews, 28, 273–283. doi: 10.1016/j.neubiorev.2004.03.004 Barnes, J., Bartlett, J. W., van de Pol, L. A., Loy, C. T., Scahill, R. I., Frost, C.,…Fox, N. C. (2009). A meta-analysis of hippocampal atrophy rates in Alzheimer's disease. Neurobiology of Aging, 30, 1711–1723. doi: 10.1016/j.neurobiolaging.2008.01.010 Bassuk, S. S., Glass, T. A., & Berkman, L. F. (1999). Social disengagement and incident cognitive decline in community-dwelling elderly persons. Annals of Internal Medicine, 131, 165–173. doi: 10.7326/0003–4819–131–3–199908030–00002

Beach, S. R., Brody, G. H., Todorov, A. A., Gunter, T. D., & Philibert, R. A. (2010). Methylation at SLC6A4 is linked to family history of child abuse: An examination of the Iowa Adoptee sample. American Journal of Medical Genetics. Part B: Neuropsychiatric Genetics, 153B, 710–703. doi: 10.1002/ajmg.b.31028 Beckie, T. M. (2012). A systematic review of allostatic load, health, and health disparities. Biological Research in Nursing, 14, 311–346. doi: 10.1177/1099800412455688 Bellwood, P. S. (1987). The Polynesians: Prehistory of an island people. London, UK: Thames and Hudson. Benguira, S., Leblond, V. S., Weber, J. P., & Hontela, A. (2002). Loss of capacity to elevate plasma cortisol in rainbow trout (Oncorhynchus mykiss) treated with a single injection of o,p '-dichlorodiphenyldichloroethane. Environmental Toxicology and Chemistry, 21, 1753–1756. Bennett, D. A., Wilson, R. S., Schneider, J. A., Evans, D. A., Mendes de Leon, C. F., Arnold, S. E.,…Bienias, J. L. (2003). Education modifies the relation of AD pathology to level of cognitive function in older persons. Neurology, 60, 1909–1915. doi: 10.1212/01.WNL.0000069923.64550.9F Berkman, L. F., & Glass, T. A. (2000). Social integration, social networks, social support, and health. In L. F. Berkman & I. Kawachi (Eds.), Social epidemiology (pp. 137–173). New York, NY: Oxford University Press. Bindon, J. R., Knight, A., Dressler, W. W., & Crews, D. E. (1997). Social context and psychosocial influences on blood pressure among American Samoans. American Journal of Physical Anthropology, 103, 7–18. doi: 10.1002/(SICI)1096–8644(199705)103:13.0.CO;2-U Bizik, G., Picard, M., Nijjar, R., Tourjman, V., McEwen, B., Lupien, S., & Juster, R. P. (2013). Allostatic load as a tool for monitoring physiological dysregulations and comorbidities in patients afflicted by severe mental illnesses. Harvard Review of Psychiatry, 21, 296–313. doi: 10.1097/HRP.0000000000000012 Blanchet, C., & Rochette, L. (2008). Nutrition and food consumption among the Inuit of Nunavik. Nunavik Inuit Health Survey 2004. http://www.inspq.qc.ca. Boldrini, M., Hen, R., Underwood, M. D., Rosoklija, G. B., Dwork, A. J., Mann, J. J., & Arango, V. (2012). Hippocampal angiogenesis and progenitor cell proliferation are increased with antidepressant use in major depression. Biological Psychiatry, 72, 562–571. doi: 10.1016/j.biopsych.2012.04.024 Boldrini, M., Santiago, A. N., Hen, R., Dwork, A. J., Rosoklija, G. B., Tamir, H.,…Mann, J. J. (2013). Hippocampal granule neuron number and dentate gyrus volume in antidepressanttreated and untreated major depression. Neuropsychopharmacology, 38, 1068–1077. doi: 10.1038/npp.2013.5

Boldrini, M., Underwood, M. D., Hen, R., Rosoklija, G. B., Dwork, A. J., John Mann, J., & Arango, V. (2009). Antidepressants increase neural progenitor cells in the human hippocampus. Neuropsychopharmacology, 34, 2376–2389. doi: 10.1038/npp.2009.75 Bonaiuto, S., Rocca, W. A., Lippi, A., Giannandrea, E., Mele, M., Cavarzeran, F., & Amaducci, L. (1995). Education and occupation as risk factors for dementia: A populationbased case-control study. Neuroepidemiology, 14, 101–109. doi: 10.1159/000109785 Borghol, N., Suderman, M., McArdle, W., Racine, A., Hallett, M., Pembrey, M.,…Szyf, M. (2012). Associations with early-life socio-economic position in adult DNA methylation. International Journal of Epidemiology, 41, 62–74. doi: 10.1093/ije/dyr147 Bosch, N. M., Riese, H., Reijneveld, S. A., Bakker, M. P., Verhulst, F. C., Ormel, J., & Oldehinkel, A. J. (2012). Timing matters: Long term effects of adversities from prenatal period up to adolescence on adolescents' cortisol stress response. The TRAILS study. Psychoneuroendocrinology, 37, 1439–1447. doi: 10.1016/j.psyneuen.2012.01.013 Boucher, O., Jacobson, S. W., Plusquellec, P., Dewailly, E., Ayotte, P., Forget-Dubois, N.,… Muckle, G. (2012). Prenatal methylmercury, postnatal lead exposure, and evidence of attention deficit/hyperactivity disorder among Inuit children in Arctic Quebec. Environmental Health Perspectives, 120, 1456–1461. doi: 10.1289/ehp.1204976 Brissette, I., Cohen, S., & Seeman, T. E. (2000). Measuring social integration and social networks. In S. Cohen, L. Underwood & B. Gottlieb (Eds.), Measuring and intervening in social support (pp. 53–85). New York, NY: Oxford University Press. Brody, G. H., Yu, T., Chen, Y. F., Kogan, S. M., Evans, G. W., Beach, S. R.,…Philibert, R. A. (2013a). Cumulative socioeconomic status risk, allostatic load, and adjustment: A prospective latent profile analysis with contextual and genetic protective factors. Developmental Psychology, 49, 913–927. doi: 10.1037/a0028847 Brody, G. H., Yu, T., Chen, E., Miller, G. E., Kogan, S. M., & Beach, S. R. (2013b). Is resilience only skin deep?: Rural African Americans' socioeconomic status–related risk and competence in preadolescence and psychological adjustment and allostatic load at age 19. Psychological Science, 24, 1285–1293. doi: 10.1177/0956797612471954 Brody, G. H., Yu, T., Chen, Y. F., Kogan, S. M., Evans, G. W., Windle, M.,…Philibert, R. A. (2013c). Supportive family environments, genes that confer sensitivity, and allostatic load among rural African American emerging adults: A prospective analysis. Journal of Family Psychology, 27, 22–29. doi: 10.1037/a0027829 Brown, A., Scales, U., Beever, W., Rickards, B., Rowley, K., & O'Dea, K. (2012). Exploring the expression of depression and distress in aboriginal men in central Australia: A qualitative study. BMC Psychiatry, 12, 97. doi: 10.1186/1471–244X-12–97 Bruce, J., Fisher, P. A., Pears, K. C., & Levine, S. (2009). Morning cortisol levels in

preschool-aged foster children: Differential effects of maltreatment type. Developmental Psychobiology, 51, 14–23. doi: 10.1002/dev.20333 Brummelte, S., & Galea, L. A. (2010). Chronic high corticosterone reduces neurogenesis in the dentate gyrus of adult male and female rats. Neuroscience, 168, 680–690. doi: 10.1016/j.neuroscience.2010.04.023 Buckley, D. I., Fu, R., Freeman, M., Rogers, K., & Helfand, M. (2009). C-reactive protein as a risk factor for coronary heart disease: A systematic review and meta-analyses for the U.S. Preventive Services Task Force. Annals of Internal Medicine, 151, 483–495. doi: 10.7326/0003–4819–151–7–200910060–00009 Burns, J. M., Cronk, B. B., Anderson, H. S., Donnelly, J. E., Thomas, G. P., Harsha, A.,… Swerdlow, R. H. (2008). Cardiorespiratory fitness and brain atrophy in early Alzheimer disease. Neurology, 71, 210–206. doi: 10.1212/01.wnl.0000317094.86209.cb Carlson, M. C., Erickson, K. I., Kramer, A. F., Voss, M. W., Bolea, N., Mielke, M.,…Fried, L. P. (2009). Evidence for neurocognitive plasticity in at-risk older adults: the experience corps program. Journals of Gerontology. Series A, Biological Sciences and Medical Sciences, 64, 1275–1282. doi: 10.1093/gerona/glp117 Carpenter, L. L., Carvalho, J. P., Tyrka, A. R., Wier, L. M., Mello, A. F., Mello, M. F.,…Price, L. H. (2007). Decreased adrenocorticotropic hormone and cortisol responses to stress in healthy adults reporting significant childhood maltreatment. Biological Psychiatry, 62, 1080– 1087. doi: 10.1016/j.biopsych.2007.05.002 Carpenter, L. L., Gawuga, C. E., Tyrka, A. R., Lee, J. K., Anderson, G. M., & Price, L. H. (2010). Association between plasma IL-6 response to acute stress and early-life adversity in healthy adults. Neuropsychopharmacology, 35, 2617–2623. doi: 10.1038/npp.2010.159 Central Intelligence Agency. (2012). American Samoa. The World Fact Book 2012. Washington, DC. https://www.cia.gov/library/publications/the-world-factbook/geos/aq.html Chan, S. R., & Blackburn, E. H. (2004). Telomeres and telomerase. Philosophical Transactions of the Royal Society of London, UK. Series B, Biological Sciences, 359, 109– 121. doi: 10.1098/rstb.2003.1370 Chang, V. W., Hillier, A. E., & Mehta, N. K. (2009). Neighborhood racial isolation, disorder and obesity. Social Forces, 87, 2063–2092. doi: 10.1353/sof.0.0188 Chapman, B. P., Khan, A., Harper, M., Stockman, D., Fiscella, K., Walton, J.,…Moynihan, J. (2009). Gender, race/ethnicity, personality, and interleukin-6 in urban primary care patients. Brain, Behavior, and Immunity, 23, 636–642. doi: 10.1016/j.bbi.2008.12.009 Chaudoir, S. R., & Fisher, J. D. (2010). The disclosure processes model: Understanding disclosure decision making and postdisclosure outcomes among people living with a concealable stigmatized identity. Psychological Bulletin, 136, 236–256. doi:

10.1037/a0018193 Chen, E., & Miller, G. E. (2012). “Shift-and-persist” strategies: Why being low in socioeconomic status isn't always bad for health. Perspectives in Psychological Science, 7, 135–158. doi: 10.1177/1745691612436694 Chen, E., Miller, G. E., Lachman, M. E., Gruenewald, T. L., & Seeman, T. E. (2012). Protective factors for adults from low-childhood socioeconomic circumstances: The benefits of shift-and-persist for allostatic load. Psychosomatic Medicine, 74, 178–186. doi: 10.1097/PSY.0b013e31824206fd Chesterman, J., & Galligan, R. (1997). Citizens without rights: Aborigines and Australian citizenship. Cambridge, UK: Cambridge University Press. Chodosh, J., Reuben, D. B., Albert, M. S., & Seeman, T. E. (2002). Predicting cognitive impairment in high-functioning community-dwelling older persons: MacArthur Studies of Successful Aging. Journal of the American Geriatric Society, 50, 1051–1060. doi: 10.1046/j.1532–5415.2002.50260.x Chollet, F., Tardy, J., Albucher, J. F., Thalamas, C., Berard, E., Lamy, C.,…Loubinoux, I. (2011). Fluoxetine for motor recovery after acute ischaemic stroke (FLAME): A randomised placebo-controlled trial. Lancet Neurology, 10, 123–130. doi: 10.1016/S1474– 4422(10)70314–8 Chow, C., Epp, J. R., Lieblich, S. E., Barha, C. K., & Galea, L. A. (2012). Sex differences in neurogenesis and activation of new neurons in response to spatial learning and memory. Psychoneuroendocrinology, 38, 1236–1250. doi: 10.1016/j.psyneuen.2012.11.007 Cicchetti, D. (2013). Annual research review: Resilient functioning in maltreated children— past, present, and future perspectives. Journal of Child Psychology and Psychiatry, 54, 402– 422. doi: 10.1111/j.1469–7610.2012.02608.x Cicchetti, D., & Garmezy, N. (1993). Prospects and promises in the study of resilience. Development and Psychopathology, 5, 497–502. doi: 10.1017/S0954579400006118 Cicchetti, D., & Rogosch, F. A. (2001). Diverse patterns of neuroendocrine activity in maltreated children. Development and Psychopathology, 13, 677–693. Cicchetti, D., Rogosch, F. A., Gunnar, M. R., & Toth, S. L. (2010). The differential impacts of early physical and sexual abuse and internalizing problems on daytime cortisol rhythm in school-aged children. Child Development, 81, 252–269. doi: 10.1111/j.1467– 8624.2009.01393.x Cicchetti, D., & Toth, S. L. (2009). The past achievements and future promises of developmental psychopathology: The coming of age of a discipline. Journal of Child Psychology and Psychiatry, 50, 16–25. doi: 10.1111/j.1469–7610.2008.01979.x

Clark, R., Anderson, N. B., Clark, V. R., & Williams, D. R. (1999). Racism as a stressor for African Americans. A biopsychosocial model. American Psychologist, 54, 805–816. doi: 10.1037/0003–066X.54.10.805 Cohen, S. (2004). Social relationships and health. American Psychologist, 59, 646–684. doi: 10.1037/0003–066X.59.8.676 Cohen, S., Gottilieb, B., & Underwood, L. (2000). Social relationships and health. In S. Cohen, L. Underwood, & B. Gottlieb (Eds.), Measuring and intervening in social support (pp. 3–25). New York, NY: Oxford University Press. Cohen, S., Schwartz, J. E., Epel, E., Kirschbaum, C., Sidney, S., & Seeman, T. (2006). Socioeconomic status, race, and diurnal cortisol decline in the Coronary Artery Risk Development in Young Adults (CARDIA) Study. Psychosomatic Medicine, 68, 41–50. doi: 10.1097/01.psy.0000195967.51768.ea Cohen, S., & Wills, T. A. (1985). Stress, social support, and the buffering hypothesis. Psychological Bulletin, 98, 310–357. doi: 10.1037/0033–2909.98.2.310 Colcombe, S. J., Erickson, K. I., Scalf, P. E., Kim, J. S., Prakash, R., McAuley, E.,…Kramer, A. F. (2006). Aerobic exercise training increases brain volume in aging humans. Journals of Gerontology: Series A, Biological Sciences and Medical Sciences, 61, 1166–1170. Colcombe, S. J., Kramer, A. F., Erickson, K. I., Scalf, P., McAuley, E., Cohen, N. J.,…Elavsky, S. (2004). Cardiovascular fitness, cortical plasticity, and aging. Proceedings of the National Academy of Sciences of the USA, 101, 3316–3321. doi: 10.1073/pnas.0400266101 Cole, S. W., Kemeny, M. E., Taylor, S. E., Visscher, B. R., & Fahey, J. L. (1996a). Accelerated course of human immunodeficiency virus infection in gay men who conceal their homosexual identity. Psychosomatic Medicine, 58, 219–231. Cole, S. W., Kemeny, M. E., Taylor, S. E., & Visscher, B. R. (1996b). Elevated physical health risk among gay men who conceal their homosexual identity. Health Psychology, 15, 243–251. Costello, E. J., Compton, S. N., Keeler, G., & Angold, A. (2003). Relationships between poverty and psychopathology: A natural experiment. Journal of the Amercan Medical Association, 290, 2023–2029. doi: 10.1001/jama.290.15.2023 Crews, D. E. (1989). Multivariate prediction of total and cardiovascular mortality in an obese Polynesian population. American Journal of Public Health, 79, 982–986. Crews, D. E. (2007). Composite estimates of physiological stress, age, and diabetes in American Samoans. American Journal of Physical Anthropology, 133, 1028–1034. doi: 10.1002/ajpa.20612 Crews, D. E., Harada, H., Aoyagi, K., Maeda, T., Alfarano, A., Sone, Y., & Kusano, Y. (2012). A pilot study of allostatic load among elderly Japanese living on Hizen-Oshima Island.

Journal of Physiological Anthropology, 31, 18. doi: 10.1186/1880–6805–31–18 Crimmins, E. M., Johnston, M., Hayward, M., & Seeman, T. (2003). Age differences in allostatic load: an index of physiological dysregulation. Experimental Gerontology, 38, 731– 734. doi: 10.1016/S0531–5565(03)00099–8 Crimmins, E. M., Kim, J. K., & Seeman, T. E. (2009). Poverty and biological risk: The earlier “aging” of the poor. Journals of Gerontology: Series A, Biological Sciences and Medical Sciences, 64, 286–292. doi: 10.1093/gerona/gln010 Crispim, D., Canani, L. H., Gross, J. L., Tschiedel, B., Souto, K. E., & Roisenberg, I. (2006). The European-specific mitochondrial cluster J/T could confer an increased risk of insulinresistance and type 2 diabetes: An analysis of the m.4216T > C and m.4917A > G variants. Annals of Human Genetics, 70, 488–495. doi: 10.1111/j.1469–1809.2005.00249.x Crowe, M., Andel, R., Pedersen, N. L., Johansson, B., & Gatz, M. (2003). Does participation in leisure activities lead to reduced risk of Alzheimer's disease? A prospective study of Swedish twins. Journals of Gerontology: Series B, Psychological Sciences and Social Sciences, 58, 249–255. Cullum, S., Huppert, F. A., McGee, M., Dening, T., Ahmed, A., Paykel, E. S., & Brayne, C. (2000). Decline across different domains of cognitive function in normal ageing: Results of a longitudinal population-based study using CAMCOG. International Journal of Geriatric Psychiatry, 15, 853–862. doi: 10.1002/1099–1166(200009)15:93.0.CO;2-T [pii] Cunningham, J., & Paradies, Y. C. (2012). Socio-demographic factors and psychological distress in Indigenous and non-Indigenous Australian adults aged 18–64 years: Analysis of national survey data. BMC Public Health, 12, 95. doi: 10.1186/1471–2458–12–95 Czubak, A., Nowakowska, E., Kus, K., Burda, K., Metelska, J., Baer-Dubowska, W., & Cichocki, M. (2009). Influences of chronic venlafaxine, olanzapine and nicotine on the hippocampal and cortical concentrations of brain-derived neurotrophic factor (BDNF). Pharmacological Reports, 61, 1017–1023. Dagyte, G., Crescente, I., Postema, F., Seguin, L., Gabriel, C., Mocaer, E.,…Koolhaas, J. M. (2011). Agomelatine reverses the decrease in hippocampal cell survival induced by chronic mild stress. Behavior and Brain Research, 218, 121–128. doi: 10.1016/j.bbr.2010.11.045 Dahl, R. E., & Gunnar, M. R. (2009). Heightened stress responsiveness and emotional reactivity during pubertal maturation: implications for psychopathology. Development and Psychopathology, 21, 1–6. doi: 10.1017/S0954579409000017 Dallaire, F., Dewailly, E., Vezina, C., Muckle, G., Weber, J. P., Bruneau, S., & Ayotte, P. (2006). Effect of prenatal exposure to polychlorinated biphenyls on incidence of acute respiratory infections in preschool Inuit children. Environmental Health Perspectives, 114,

1301–1305. doi: 10.1289/ehp.8683 Danese, A., Caspi, A., Williams, B., Ambler, A., Sugden, K., Jadwiga, M.,…Arseneault, L. (2011). Biological embedding of stress through inflammation processes in childhood. Molecular Psychiatry, 16, 244–246. doi: 10.1038/mp.2010.5 Danese, A., & McEwen, B. S. (2012). Adverse childhood experiences, allostasis, allostatic load, and age-related disease. Physiology and Behavior, 106, 29–39. doi: 10.1016/j.physbeh.2011.08.019 Danese, A., Moffitt, T. E., Harrington, H., Milne, B. J., Polanczyk, G., Pariante, C. M.,… Caspi, A. (2009). Adverse childhood experiences and adult risk factors for age-related disease: Depression, inflammation, and clustering of metabolic risk markers. Archives of Pediatrics and Adolescent Medicine, 163, 1135–1143. doi: 10.1001/archpediatrics.2009.214 Daniel, M., O'Dea, K., Rowley, K. G., McDermott, R., & Kelly, S. (1999). Glycated hemoglobin as an indicator of social environmental stress among indigenous versus westernized populations. Preventive Medicine, 29, 405–413. doi: 10.1006/pmed.1999.0559 Daniels, N., Kennedy, B., & Kawachi, I. (2000). Justice is good for our health. In J. Cohen & J. Rogers (Eds.), Is inequality bad for our health?—New Democracy Forum Series (p. 99). Boston, MA: Beacon Press. Dauvergne, M., Scrim, K., & Brennan, S. (2008). Hate crime in Canada, 2006. Statistics Canada. Canadian Centre for Justice Statistics. Ottawa. Davis, L. L., Weaver, M., Zamrini, E., Stevens, A., Kang, D. H., & Parker, C. R., Jr. (2004). Biopsychological markers of distress in informal caregivers. Biological Research in Nursing, 6, 90–99. doi: 10.1177/1099800404267353 De Benedictis, G., Rose, G., Carrieri, G., De Luca, M., Falcone, E., Passarino, G.,… Franceschi, C. (1999). Mitochondrial DNA inherited variants are associated with successful aging and longevity in humans. Federation of American Societies for Experimental Biology Journal, 13, 1532–1536. Decker, S. A. (2000). Salivary cortisol and social status among Dominican men. Hormones and Behavior, 38, 29–38. doi: 10.1006/hbeh.2000.1597 Dedovic, K., Wadiwalla, M., Engert, V., & Pruessner, J. C. (2009). The role of sex and gender socialization in stress reactivity. Developmental Psychology, 45, 45–55. doi: 10.1037/a0014433 Demakakos, P., Nazroo, J., Breeze, E., & Marmot, M. (2008). Socioeconomic status and health: the role of subjective social status. Social Science and Medicine, 67, 330–340. doi: 10.1016/j.socscimed.2008.03.038 Devlin, A. M., Brain, U., Austin, J., & Oberlander, T. F. (2010). Prenatal exposure to maternal

depressed mood and the MTHFR C677T variant affect SLC6A4 methylation in infants at birth. PLoS One, 5, e12201. doi: 10.1371/journal.pone.0012201 De Vogli, R., Ferrie, J. E., Chandola, T., Kivimaki, M., & Marmot, M. G. (2007). Unfairness and health: evidence from the Whitehall II Study. Journal of Epidemiology and Community Health, 61, 513–518. doi: 10.1136/jech.2006.052563 de Vugt, M. E., Nicolson, N. A., Aalten, P., Lousberg, R., Jolle, J., & Verhey, F. R. (2005). Behavioral problems in dementia patients and salivary cortisol patterns in caregivers. Journal of Neuropsychiatry and Clinical Neurosciences, 17, 201–207. doi: 10.1176/appi.neuropsych.17.2.201 DeWailly, P., Ayotte, P., Pereg, P., Dery, S., Dallaire, R., Fontaine, J., & Côté, S. (2007a). Exposure to environmental contaminants in Nunavik: Metals. Nunavik Inuit Health Survey 2004. DeWailly, P., Dallaire, R., Pereg, P., Ayotte, P., Fontaine, J., & Dery, S. (2007b). Exposure to environmental contaminants in Nunavik: Persistent organic pollutants and new contaminants of concern. Nunavik Inuit Health Survey 2004. Dhabhar, F. S., Malarkey, W. B., Neri, E., & McEwen, B. S. (2012). Stress-induced redistribution of immune cells—from barracks to boulevards to battlefields: A tale of three hormones—Curt Richter Award winner. Psychoneuroendocrinology, 37, 1345–1368. doi: 10.1016/j.psyneuen.2012.05.008 Dickerson, S. S., & Kemeny, M. E. (2004). Acute stressors and cortisol responses: A theoretical integration and synthesis of laboratory research. Psychological Bulletin, 130, 355– 391. doi: 10.1037/0033–2909.130.3.355 DiMauro, S., & Schon, E. A. (2003). Mitochondrial respiratory-chain diseases. New England Journal of Medicine, 348, 2656–2668. doi: 10.1056/NEJMra022567 Dion, K. L. (2001). The social psychology of perceived prejudice and discrimination. Canadian Psychology, 43, 1–10. Dowd, J. B., Simanek, A. M., & Aiello, A. E. (2009). Socio-economic status, cortisol and allostatic load: A review of the literature. International Journal of Epidemiology, 38, 1297– 1309. doi: 10.1093/ije/dyp277 Dressler, W. W., Balieiro, M. C., & Dos Santos, J. E. (1997). The cultural construction of social support in Brazil: Associations with health outcomes. Culture, Medicine and Psychiatry 21, 303–335. Duru, O. K., Harawa, N. T., Kermah, D., & Norris, K. C. (2012). Allostatic load burden and racial disparities in mortality. Journal of the National Medical Association, 104, 89–95. Eisenberger, N. I., Taylor, S. E., Gable, S. L., Hilmert, C. J., & Lieberman, M. D. (2007).

Neural pathways link social support to attenuated neuroendocrine stress responses. Neuroimage, 35, 1601–1612. doi: 10.1016/j.neuroimage.2007.01.038 Elizalde, N., Garcia-Garcia, A. L., Totterdell, S., Gendive, N., Venzala, E., Ramirez, M. J.,… Tordera, R. M. (2010). Sustained stress-induced changes in mice as a model for chronic depression. Psychopharmacology (Berlin) 210, 393–406. doi: 10.1007/s00213–010–1835–6 Engert, V., Efanov, S. I., Dedovic, K., Dagher, A., & Pruessner, J. C. (2011). Increased cortisol awakening response and afternoon/evening cortisol output in healthy young adults with low early life parental care. Psychopharmacology (Berlin), 214, 261–268. doi: 10.1007/s00213– 010–1918–4 Epel, E. S. (2009). Psychological and metabolic stress: a recipe for accelerated cellular aging? Hormones (Athens), 8, 7–22. Epel, E. S., Blackburn, E. H., Lin, J., Dhabhar, F. S., Adler, N. E., Morrow, J. D., & Cawthon, R. M. (2004). Accelerated telomere shortening in response to life stress. Proceedings of the National Academy of Sciences of the USA, 101, 17312–17315. doi: 10.1073/pnas.0407162101 Erickson, K. I., Prakash, R. S., Voss, M. W., Chaddock, L., Hu, L., Morris, K. S.,…Kramer, A. F. (2009). Aerobic fitness is associated with hippocampal volume in elderly humans. Hippocampus, 19, 1030–1039. doi: 10.1002/hipo.20547 Erickson, K. I., Voss, M. W., Prakash, R. S., Basak, C., Szabo, A., Chaddock, L.,…Kramer, A. F. (2011). Exercise training increases size of hippocampus and improves memory. Proceedings of the National Academy of Sciences of the USA, 108, 3017–3022. doi: 10.1073/pnas.1015950108 Ernst, C., Marshall, C. R., Shen, Y., Metcalfe, K., Rosenfeld, J., Hodge, J. C.,…Talkowski, M. E. (2012). Highly penetrant alterations of a critical region including BDNF in human psychopathology and obesity. Archives of General Psychiatry, 69, 1238–1246. doi: 10.1001/archgenpsychiatry.2012.660 Ernst, C., Wanner, B., Brezo, J., Vitaro, F., Tremblay, R., & Turecki, G. (2011). A deletion in tropomyosin-related kinase B and the development of human anxiety. Biological Psychiatry, 69, 604–607. doi: 10.1016/j.biopsych.2010.10.008 Essex, M. J., Thomas Boyce, W., Hertzman, C., Lam, L. L., Armstrong, J. M., Neumann, S. M., & Kobor, M. S. (2011). Epigenetic vestiges of early developmental adversity: Childhood stress exposure and DNA methylation in adolescence. Child Development, 84, 58–75. doi: 10.1111/j.1467–8624.2011.01641.x Evans, G. W., Lepore, S. J., Shejwal, B. R., & Palsane, M. N. (1998). Chronic residential crowding and children's well-being: an ecological perspective. Child Development, 69, 1514–1523.

Evans, O., & Steptoe, A. (2002). The contribution of gender-role orientation, work factors and home stressors to psychological well-being and sickness absence in male- and femaledominated occupational groups. Social Science and Medicine, 54, 481–492. doi: 10.1016/S0277–9536(01)00044–2 Fagundes, C. P., Glaser, R., & Kiecolt-Glaser, J. K. (2013). Stressful early life experiences and immune dysregulation across the lifespan. Brain, Behavior, and Immunity, 27, 8–12. doi: 10.1016/j.bbi.2012.06.014 Fanselow, M. S., & Dong, H. W. (2010). Are the dorsal and ventral hippocampus functionally distinct structures? Neuron, 65, 7–19. doi: 10.1016/j.neuron.2009.11.031 Felitti, V. J., Anda, R. F., Nordenberg, D., Williamson, D. F., Spitz, A. M., Edwards, V.,… Marks, J. S. (1998). Relationship of childhood abuse and household dysfunction to many of the leading causes of death in adults. The Adverse Childhood Experience (ACE) Study. American Journal of Preventive Medicine, 14, 245–258. doi: 10.1186/1471–2458–9–106 Fernandez-Moreno, M., Tamayo, M., Soto-Hermida, A., Mosquera, A., Oreiro, N., FernandezLopez, C.,…Blanco, F. J. (2011). mtDNA haplogroup J modulates telomere length and nitric oxide production. BMC Musculoskeletal Disorders, 12, 283. doi: 10.1186/1471–2474–12– 283 Ferreira, C. L., Santos, L. M. O., & Maia, E. M. C. (2012). Resilience among the elderly cared for by the Primary Healthcare Network in a city of northeast Brazil. Revista da Escola de Enfermagem da USP, 46, 328–334. Fisher, P. A., Gunnar, M. R., Dozier, M., Bruce, J., & Pears, K. C. (2006). Effects of therapeutic interventions for foster children on behavioral problems, caregiver attachment, and stress regulatory neural systems. Annals of the New York Academy of Science, 1094, 215–225. doi: 10.1196/annals.1376.023. Flinn, M. V., & England, B. G. (1997). Social economics of childhood glucocorticoid stress response and health. American Journal of Physical Anthropology, 102, 33–53. Fortes, T. F. R., Portugues, M. W., & Argimon, I. I. L. (2009). Resilience in the elderly and its relationship with socio-demographic variables and cognitive functions. Estudos de Psicologia, 26, 455–463. doi: 10.1590/S0103–166X2009000400006 Fortin, M. C., Cory-Slechta, D. A., Ohman-Strickland, P., Nwankwo, C., Yanger, T. S., Todd, A. C.,…Fiedler, N. (2012). Increased lead biomarker levels are associated with changes in hormonal response to stress in occupationally exposed male participants. Environmental Health Perspectives, 120, 278–283. doi: 10.1289/ehp.1103873 Fratiglioni, L., Paillard-Borg, S., & Winblad, B. (2004). An active and socially integrated lifestyle in late life might protect against dementia. Lancet Neurology, 3, 343–353. doi: 10.1016/S1474–4422(04)00767–7

Fratiglioni, L., Wang, H. X., Ericsson, K., Maytan, M., & Winblad, B. (2000). Influence of social network on occurrence of dementia: A community-based longitudinal study. Lancet, 355, 1315–1319. doi: 10.1016/S0140–6736(00)02113–9 Friedland, R. P., Fritsch, T., Smyth, K. A., Koss, E., Lerner, A. J., Chen, C. H.,…Debanne, S. M. (2001). Patients with Alzheimer's disease have reduced activities in midlife compared with healthy control-group members. Proceedings of the National Academy of Sciences of the USA, 98, 3440–2445. doi: 10.1073/pnas.061002998 Friedman, E. M., Williams, D. R., Singer, B. H., & Ryff, C. D. (2009). Chronic discrimination predicts higher circulating levels of E-selectin in a national sample: The MIDUS study. Brain, Behavior, and Immunity, 23, 684–692. doi: 10.1016/j.bbi.2009.01.002 Fries, A. B., Ziegler, T. E., Kurian, J. R., Jacoris, S., & Pollak, S. D. (2005). Early experience in humans is associated with changes in neuropeptides critical for regulating social behavior. Proceedings of the National Academy of Sciences of the USA, 102, 17237–17240. doi: 10.1073/pnas.0504767102 Fries, J. H., Hellhammer, J., & Hellhammer, D. H. (2005). A new view of hypocortisolism. Psychoneuroendocrinology, 30, 1010–1016. doi: 10.1016/j.psyneuen.2005.04.006 Fukagawa, N. K., Li, M., Liang, P., Russell, J. C., Sobel, B. E., & Absher, P. M. (1999). Aging and high concentrations of glucose potentiate injury to mitochondrial DNA. Free Radical Biology and Medicine, 27, 1437–1443. doi: 10.1016/S0891–5849(99)00189–6 Fuller-Rowell, T. E., Doan, S. N., & Eccles, J. S. (2012). Differential effects of perceived discrimination on the diurnal cortisol rhythm of African Americans and Whites. Psychoneuroendocrinology, 37, 107–118. doi: 10.1016/j.psyneuen.2011.05.011 Fuller-Rowell, T. E., Evans, G. W., & Ong, A. D. (2012). Poverty and health: The mediating role of perceived discrimination. Psychological Science, 23, 734–739. doi: 10.1177/0956797612439720 Gaiolli, C. C. L., Furegato, A. F., & Santos, J. L. F. (2012). Profile of caregivers of elderly people with Alzheimer's disease associated to resilience. Texto and Contexto Enfermagem, 21, 150–157. doi: 10.1590/S0104–07072012000100017 Gallo, L. C., Penedo, F. J., Espinosa de los Monteros, K., & Arguelles, W. (2009). Resiliency in the face of disadvantage: Do Hispanic cultural characteristics protect health outcomes? Journal of Personality, 77, 1707–1746. doi: 10.1111/j.1467–6494.2009.00598.x Gardner, A., & Boles, R. G. (2011). Beyond the serotonin hypothesis: Mitochondria, inflammation and neurodegeneration in major depression and affective spectrum disorders. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 35, 730–743. doi: 10.1016/j.pnpbp.2010.07.030 Gates, G. J. (2010). Sexual minorities in the 2008 General Social Survey: Coming Out and

demographic characteristics. Los Angeles, CA: Williams Institute. Ge, S., Yang, C. H., Hsu, K. S., Ming, G. L., & Song, H. (2007). A critical period for enhanced synaptic plasticity in newly generated neurons of the adult brain. Neuron, 54, 559–566. doi: 10.1016/j.neuron.2007.05.002 Gendron, A. D., Bishop, C. A., Fortin, R., & Hontela, A. (1997). In vivo testing of the functional integrity of the corticosterone-producing axis in mudpuppy (amphibia) exposed to chlorinated hydrocarbons in the wild. Environmental Toxicology and Chemistry, 16, 1694– 1706. Geronimus, A. T., Hicken, M., Keene, D., & Bound, J. (2006). “Weathering” and age patterns of allostatic load scores among Blacks and Whites in the United States. American Journal of Public Health 96, 826–833. doi: 10.2105/AJPH.2004.060749 Gersner, R., Toth, E., Isserles, M., & Zangen, A. (2010). Site-specific antidepressant effects of repeated subconvulsive electrical stimulation: Potential role of brain-derived neurotrophic factor. Biological Psychiatry, 67, 125–132. doi: 10.1016/j.biopsych.2009.09.015 Ghiso, J., & Frangione, B. (2002). Amyloidosis and Alzheimer's disease. Advanced Drug Delivery Reviews, 54, 1539–1551. Gilbert, R., Widom, C. S., Browne, K., Fergusson, D., Webb, E., & Janson, S. (2009). Burden and consequences of child maltreatment in high-income countries. Lancet, 373, 68–81. doi: 10.1016/S0140–6736(08)61706–7 Glei, D. A., Goldman, N., Shkolnikov, V. M., Jdanov, D., Shkolnikova, M., Vaupel, J. W., & Weinstein, M. (2013). Perceived stress and biological risk: Is the link stronger in Russians than in Taiwanese and Americans? Stress, 16, 411–420. doi: 10.3109/10253890.2013.789015 Gomez-Duran, A., Pacheu-Grau, D., Lopez-Gallardo, E., Diez-Sanchez, C., Montoya, J., Lopez-Perez, M. J., & Ruiz-Pesini, E. (2010). Unmasking the causes of multifactorial disorders: OXPHOS differences between mitochondrial haplogroups. Human Molecular Genetics, 19, 3343–3353. doi: 10.1093/hmg/ddq246 Goodman, E., McEwen, B. S., Huang, B., Dolan, L. M., & Adler, N. E. (2005). Social inequalities in biomarkers of cardiovascular risk in adolescence. Psychosomatic Medicine, 67, 9–15. doi: 10.1097/01.psy.0000149254.36133.1a Gotmann, W., & Wingfield, J. C. (2004). Allostatic load, social status and stress hormones: The costs of social status matter. Animal Behaviour, 67, 591–602. Gouin, J. P., Glaser, R., Malarkey, W. B., Beversdorf, D., & Kiecolt-Glaser, J. K. (2012). Childhood abuse and inflammatory responses to daily stressors. Annals of Behavioral Medicine, 44, 287–292. doi: 10.1007/s12160–012–9386–1 Goymann, W., & Wingfield, J. C. (2004). Allostatic load, social status and stress hormones:

The costs of social status matter. Animal Behaviour, 67, 591–602. doi: 10.1016/j.anbehav.2003.08.007 Gracey, M., & King, M. (2009). Indigenous health part 1: Determinants and disease patterns. Lancet, 374, 65–75. doi: 10.1016/S0140–6736(09)60914–4 Green, J. G., McLaughlin, K. A., Berglund, P. A., Gruber, M. J., Sampson, N. A., Zaslavsky, A. M., & Kessler, R. C. (2010). Childhood adversities and adult psychiatric disorders in the National Comorbidity Survey replication I: Associations with first onset of DSM-IV disorders. Archives of General Psychiatry, 67, 113–123. doi: 10.1001/archgenpsychiatry.2009.186 Greene, K., Derlega, V. J., & Matthews, A. (2006). Self-disclosure in personal relationships. In A. L. Vangelisti & D. Perlman (Eds.), Cambridge handbook of personal relationships (pp. 409–427). New York, NY: Cambridge University Press. Greeson, J. M., Hurwitz, B. E., Llabre, M. M., Schneiderman, N., Penedo, F. J., & Klimas, N. G. (2008). Psychological distress, killer lymphocytes and disease severity in HIV/AIDS. Brain Behavior and Immunity, 22, 901–911. Gruenewald, T. L., Karlamangla, A. S., Hu, P., Stein-Merkin, S., Crandall, C., Koretz, B., & Seeman, T. E. (2012). History of socioeconomic disadvantage and allostatic load in later life. Social Science and Medicine, 74, 75–83. doi: 10.1016/j.socscimed.2011.09.037 Gruenewald, T. L., Seeman, T. E., Ryff, C. D., Karlamangla, A. S., & Singer, B. H. (2006). Combinations of biomarkers predictive of later life mortality. Proceedings of the National Academy of Sciences of the USA, 103, 14158–14163. doi: 10.1073/pnas.0606215103 Gump, B. B., Stewart, P., Reihman, J., Lonky, E., Darvill, T., Parsons, P. J., & Granger, D. A. (2008). Low-level prenatal and postnatal blood lead exposure and adrenocortical responses to acute stress in children. Environmental Health Perspectives, 116, 249–255. Gunnar, M. R., & Vazquez, D. M. (2001). Low cortisol and a flattening of expected daytime rhythm: potential indices of risk in human development. Development and Psychopathology, 13, 515–538. doi: 10.1017/S0954579401003145 Gunnar, M. R., & Vazquez, D. M. (2006). Stress neurobiology and developmental psychopathology. In D. Cicchetti & J. C. Donald (Eds.), Developmental psychopathology (pp. 533–577). Hoboken, NJ: Wiley. Hajat, A., Diez-Roux, A., Franklin, T. G., Seeman, T., Shrager, S., Ranjit, N.,…Kirschbaum, C. (2010). Socioeconomic and race/ethnic differences in daily salivary cortisol profiles: The Multi-Ethnic Study of Atherosclerosis. Psychoneuroendocrinology, 35, 932–943. doi: 10.1016/j.psyneuen.2009.12.009 Hamer, M. (2012). Psychosocial stress and cardiovascular disease risk: the role of physical activity. Psychosomatic Medicine, 74, 896–903. doi: 10.1097/PSY.0b013e31827457f4

Handschin, C., & Spiegelman, B. M. (2008). The role of exercise and PGC1alpha in inflammation and chronic disease. Nature, 454, 463–469. doi: 10.1038/nature07206 Hanson, N. D., Owens, M. J., & Nemeroff, C. B. (2011). Depression, antidepressants, and neurogenesis: A critical reappraisal. Neuropsychopharmacology, 36, 2589–2602. doi: 10.1038/npp.2011.220 Hara, Y., Yuk, F., Puri, R., Janssen, W. G., Rapp, P. R., & Morrison, J. H. (2014). Presynaptic mitochondrial morphology in monkey prefrontal cortex correlates with working memory and is improved with estrogen treatment. Proceedings from the National Academy of Sciences of the USA, 111, 486–491. doi: 10.1073/pnas.1311310110 Harrell, J. P., Hall, S., & Taliaferro, J. (2003). Physiological responses to racism and discrimination: an assessment of the evidence. American Journal of Public Health, 93, 243– 248. Harrell, J. P., Merrit, M. M., & Kalu, J. (1998). Racism, stress, and disease. In R. L. J. (Ed.), African American mental health, 247–280. Hampton, VA: Cobb & Henry. Harrison, G. A. (2001). Comparative stress in human socities. Journal of Physiology and Anthropology, 20, 49–53. doi: 10.2114/jpa.20.49 Hatzenbuehler, M. L. (2009). How does sexual minority stigma “get under the skin”? A psychological mediation framework. Psychological Bulletin, 135, 707–730. doi: 10.1037/a0016441 Hatzenbuehler, M. L. (2010). Social factors as determinants of mental health disparities in LGB populations: implications for public policy. Social Issues and Policy Review, 4, 31–62. Hatzenbuehler, M. L., Bellatorre, A., Lee, Y., Finch, B. K., Muennig, P., & Fiscella, K. (2014). Structural stigma and all-cause mortality in sexual minority populations. Social Science and Medicine, 103, 33–41. doi: 10.1016/j.socscimed.2013.06.005 Hatzenbuehler, M. L., Keyes, K. M., & Hasin, D. S. (2009). State-level policies and psychiatric morbidity in lesbian, gay, and bisexual populations. American Journal of Public Health, 99, 2275–2281. doi: 10.2105/AJPH.2008.153510 Hatzenbuehler, M. L., & McLaughlin, K. A. (2013). Structural Stigma and HypothalamicPituitary—Adrenocortical Axis Reactivity in Lesbian, Gay, and Bisexual Young Adults. Annals of Behavioral Medicine, 47(1), 39–47. http://doi.org/10.1007/s12160-013-9556-9 Hatzenbuehler, M. L., McLaughlin, K. A., Keyes, K. M., & Hasin, D. S. (2010). The impact of institutional discrimination on psychiatric disorders in lesbian, gay, and bisexual populations: A prospective study. American Journal of Public Health, 100, 452–459. doi: 10.2105/AJPH.2009.168815 Hatzenbuehler, M. L., O'Cleirigh, C., Mayer, K., Safren, S. A., & Bradford, J. (2012). Effects

of same-sex marriage laws on health care use and expenditures in sexual minority men: A quasi-natural experiment. American Journal of Public Health, 102, 285–291. Heim, C., Newport, D. J., Heit, S., Graham, Y. P., Wilcox, M., Bonsall, R.,…Nemeroff, C. B. (2000). Pituitary-adrenal and autonomic responses to stress in women after sexual and physical abuse in childhood. Journal of the American Medical Association, 284, 592–597. doi: 10.1001/jama.284.5.592 Hendrickson, S. L., Hutcheson, H. B., Ruiz-Pesini, E., Poole, J. C., Lautenberger, J., Sezgin, E.,…O'Brien, S. J. (2008). Mitochondrial DNA haplogroups influence AIDS progression. AIDS, 22, 2429–2439. doi: 10.1097/QAD.0b013e32831940bb Herek, G. M. (1992). The social context of hate crimes: Notes on cultural heterosexism. In G. M. Herek & K. Berrill (Eds.), Hate crimes: Confronting violence against lesbians and gay men (pp. 89–104). Thousand Oaks, CA: Sage. Herek, G. M., & Garnets, L. D. (2007). Sexual orientation and mental health. Annual Review of Clinical Psychology, 3, 353–375. Herek, G. M., Gillis, J. R., & Cogan, J. C. (1999). Psychological sequelae of hate-crime victimization among lesbian, gay, and bisexual adults. Journal of Consulting and Clinical Psychology, 67, 945–951. Herman, J. P., Ostrander, M. M., Mueller, N. K., & Figueiredo, H. (2005). Limbic system mechanisms of stress regulation: Hypothalamo-pituitary-adrenocortical axis. Progress in Neuropsychopharmacology and Biological Psychiatry, 29, 1201–1213. doi: 10.1016/j.pnpbp.2005.08.006 Hershberger, S. L., Pilkington, N. W., & D'Augelli, A. R. (1997). Predictors of suicide attempts among gay, lesbian, and bisexual youth. Journal of Adolescent Research, 12, 477– 497. Hertzman, C. (1999). The biological embedding of early experience and its effects on health in adulthood. Annals of the New York Academy of Science, 896, 85–95. doi: 10.1111/j.1749– 6632.1999.tb08107.x Heyn, P., Abreu, B. C., & Ottenbacher, K. J. (2004). The effects of exercise training on elderly persons with cognitive impairment and dementia: a meta-analysis. Archives of Physical Medicine and Rehabilitation, 85, 1694–1704. doi: 10.1016/j.apmr.2004.03.019 Hickson, D. A., Diez Roux, A. V., Gebreab, S. Y., Wyatt, S. B., Dubbert, P. M., Sarpong, D. F., …Taylor, H. A. (2012). Social patterning of cumulative biological risk by education and income among African Americans. American Journal of Public Health, 102, 1362–1269. doi: 10.2105/AJPH.2011.300444 Hirao, K., Ohnishi, T., Hirata, Y., Yamashita, F., Mori, T., Moriguchi, Y.,…Asada, T. (2005). The prediction of rapid conversion to Alzheimer's disease in mild cognitive impairment using

regional cerebral blood flow SPECT. Neuroimage, 28, 1014–1021. doi: 10.1016/j.neuroimage.2005.06.066 Hontela, A., Dumont, P., Duclos, D., & Fortin, R. (1995). Endocrine and metabolic dysfunction in yellow perch, Perca flavescens, exposed to organic contaminants and heavy-metals in the St Lawrence River. Environmental Toxicology and Chemistry, 14, 725–731. Hontela, A., Rasmussen, J. B., Audet, C., & Chevalier, G. (1992). Impaired cortisol stress response in fish from environments polluted by PAHS, PCBS, and mercury. Archives of Environmental Contamination and Toxicology, 22, 278–283. Hood, D. A. (2009). Mechanisms of exercise-induced mitochondrial biogenesis in skeletal muscle. Applied Physiology, Nutrition, and Metabolism, 34, 465–472. doi: 10.1139/h09–045 Howren, M. B., Lamkin, D. M., & Suls, J. (2009). Associations of depression with C-reactive protein, IL-1, and IL-6: A meta-analysis. Psychosomatic Medicine, 71, 171–186. doi: 10.1097/PSY.0b013e3181907c1b Hu, F. B., Willett, W. C., Li, T., Stampfer, M. J., Colditz, G. A., & Manson, J. E. (2004). Adiposity as compared with physical activity in predicting mortality among women. New England Journal of Medicine, 351, 2694–2703. doi: 10.1056/NEJMoa042135 Hu, P., Wagle, N., Goldman, N., Weinstein, M., & Seeman, T. E. (2007). The associations between socioeconomic status, allostatic load and measures of health in older Taiwanese persons: Taiwan Social Environment and Biomarkers of Aging study. Journal of Biosocial Science, 39, 545–556. doi: 10.1017/S0021932006001556 Huebner, D. M., & Davis, M. C. (2005). Gay and bisexual men who disclose their sexual orientations in the workplace have higher workday levels of salivary cortisol and negative affect. Annals of Behavioral Medicine, 30, 260–267. doi: 10.1207/s15324796abm3003_10 Huebner, D. M., Rebchook, G. M., & Kegeles, S. M. (2004). Experiences of harassment, discrimination, and physical violence among young gay and bisexual men. American Journal of Public Health, 94, 1200–1203. Hunter, E. (2013). Indicators of psychoses or psychoses as indicators: the relationship between Indigenous social disadvantage and serious mental illness. Australian Psychiatry, 21, 22–26. doi: 10.1177/1039856212460598 Hunter, E. M., Gynther, B. D., Anderson, C. J., Onnis, L. A., Nelson, J. R.,…Groves, A. R. (2012). Psychosis in Indigenous populations of Cape York and the Torres Strait. Medical Journal of Australia, 196, 133–135. doi: 10.5694/mja11.10118 Ibrahim, M. M., & Damasceno, A. (2012). Hypertension in developing countries. Lancet, 380, 611–619. doi: 10.1016/S0140–6736(12)60861–7 Institute for Applied Economic Research. (2009). Workload: Evolution and main changes in

the country, 1–11. http://ipea.gov.br/agencia/images/stories/PDFs/comunicado/090729_comunicadoipea24.pdf Institute of Medicine. (2011). The health of lesbian, gay, bisexual, and transgender people: Building a foundation for better understanding. Washington, DC: National Academy Press. Ivannikov, M. V., Sugimori, M., & Llinas, R. R. (2013). Synaptic vesicle exocytosis in hippocampal synaptosomes correlates directly with total mitochondrial volume. Journal of Molecular Neuroscience, 49, 223–230. doi: 10.1007/s12031–012–9848–8 Jackman, M. R. (1994). The velvet glove: Paternalism and conflict in gender, class, and race relations. Los Angeles, CA: University of California Press. Jackson, J. S., Knight, K. M., & Rafferty, J. A. (2010). Race and unhealthy behaviors: Chronic stress, the HPA axis, and physical and mental health disparities over the life course. American Journal of Public Health, 100, 933–939. doi: 10.2105/AJPH.2008.143446 James, S. A. (1994). John Henryism and the health of African-Americans. Culture, Medicine, and Psychiatry, 18, 163–182. Jayatissa, M. N., Bisgaard, C., Tingstrom, A., Papp, M., & Wiborg, O. (2006). Hippocampal cytogenesis correlates to escitalopram-mediated recovery in a chronic mild stress rat model of depression. Neuropsychopharmacology, 31, 2395–2404. doi: 10.1038/sj.npp.1301041 Jayatissa, M. N., Bisgaard, C. F., West, M. J., & Wiborg, O. (2008). The number of granule cells in rat hippocampus is reduced after chronic mild stress and re-established after chronic escitalopram treatment. Neuropharmacology, 54, 530–541. doi: 10.1016/j.neuropharm.2007.11.009 Jayatissa, M. N., Henningsen, K., West, M. J., & Wiborg, O. (2009). Decreased cell proliferation in the dentate gyrus does not associate with development of anhedonic-like symptoms in rats. Brain Research, 1290, 133–141. doi: 10.1016/j.brainres.2009.07.001 Jin, K., Zhu, Y., Sun, Y., Mao, X. O., Xie, L., & Greenberg, D. A. (2002). Vascular endothelial growth factor (VEGF) stimulates neurogenesis in vitro and in vivo. Proceedings of the National Academy of Sciences of the USA, 99, 11946–11950. doi: 10.1073/pnas.182296499 Jirtle, R. L., & Skinner, M. K. (2007). Environmental epigenomics and disease susceptibility. Nature Reviews Genetics, 8, 253–262. doi: 10.1038/nrg2045 Johansson, M., Nilsson, S., & Lund, B. O. (1998). Interactions between methylsulfonyl PCBs and the glucocorticoid receptor. Environmental Health Perspectives, 106, 769–772. Johnson, J. L., & Repta, R. (2011). Sex and gender: Beyond the binaries. In J. L. Oliffe & L. Greaves (Eds.), Designing and conducting gender, sex, and health research, 17–37. Los Angeles, California: Sage. Johnson, J. L., Repta, R., & Kaylan, S. (2011). Implications of sex and gender for health

research. In J. L. Oliffe & L. Greaves (Eds.), Designing and conducting gender, sex, and health research, 39–64. Los Angeles, California: Sage Jorm, A. F., Bourchier, S. J., Cvetkovski, S., & Stewart, G. (2012). Mental health of Indigenous Australians: A review of findings from community surveys. Medical Journal of Australia, 196, 118–121. doi: 10.5694/mja11.10041 Juster, R. P., Bizik, G., Picard, M., Arsenault-Lapierre, G., Sindi, S., Trepanier, L.,…Lupien, S. J. (2011). A transdisciplinary perspective of chronic stress in relation to psychopathology throughout life span development. Development and Psychopathology, 23, 725–776. doi: 10.1017/S0954579411000289 Juster, R. P., & Lupien, S. J. (2012a). Chronic stress and allostatic load. In K. SchenckGustafsson, P. R. DeCola, D. W. Pfaff, & D. S. Pisetsky (Eds.), Handbook of clinical gender medicine (pp. 70–81). Basel, Switzerland: Karger AG. Juster, R. P., & Lupien, S. (2012b). A sex- and gender-based analysis of allostatic load and physical complaints. Gender Medicine, 9, 511–523. doi: 10.1016/j.genm.2012.10.008 Juster, R. P., & Lupien, S. J. (2012c). Sex and gender in stress research: The metamorphosis of a field. In IGH (Ed.) What a difference sex and gender make: A gender, sex and health research casebook. Vancouver, BC: Canadian Institutes for Health Research Institute of Gender and Health. Juster, R. P., McEwen, B. S., & Lupien, S. J. (2010). Allostatic load biomarkers of chronic stress and impact on health and cognition. Neuroscience and Biobehavioral Reviews, 35, 2– 16. doi: 10.1016/j.neubiorev.2009.10.002 Juster, R. P., Moskowitz, D. S., Lavoie, J., & D'Antono, B. (2013a). Sex-specific interaction effects of age, occupational status, and workplace stress on psychiatric symptoms and allostatic load among healthy Montreal workers. Stress, 16, 616–629. doi: 10.3109/10253890.2013.835395 Juster, R. P., Smith, N. G., Ouellet, E., Sindi, S., & Lupien, S. J. (2013b). Sexual orientation and disclosure in relation to psychiatric symptoms, diurnal cortisol, and allostatic load. Psychosomatic Medicine, 75, 103–116. doi: 10.1097/PSY.0b013e3182826881 Karatsoreos, I. N., & McEwen, B. S. (2011). Psychobiological allostasis: resistance, resilience and vulnerability. Trends in Cognitive Science, 15, 576–584. doi: 10.1016/j.tics.2011.10.005 Karlamangla, A. S., Singer, B. H., & Seeman, T. E. (2006). Reduction in allostatic load in older adults is associated with lower all-cause mortality risk: MacArthur studies of successful aging. Psychosomatic Medicine, 68, 500–507. doi: 10.1097/01.psy.0000221270.93985.82 Karlen, J., Ludvigsson, J., Frostell, A., Theodorsson, E., & Faresjo, T. (2011). Cortisol in hair measured in young adults—a biomarker of major life stressors? BMC Clinical Pathology, 11,

12. doi: 10.1186/1472–6890–11–12 Karp, A., Kareholt, I., Qiu, C., Bellander, T., Winblad, B., & Fratiglioni, L. (2004). Relation of education and occupation-based socioeconomic status to incident Alzheimer's disease. American Journal of Epidemiology, 159, 175–183. doi: 10.1093/aje/kwh018 Katzman, R. (1995). Can late life social or leisure activities delay the onset of dementia? Journal of the American Geriatric Society, 43, 583–584. Kawachi, I., & Kennedy, B. (2002). The health of nations: Why inequality is harmful to your health. New York, NY: New Press. Kempermann, G. (2011). Adult neurogenesis 2. New York, NY: Oxford University Press. Kemppainen, N. M., Aalto, S., Karrasch, M., Nagren, K., Savisto, N., Oikonen, V.,…Rinne, J. O. (2008). Cognitive reserve hypothesis: Pittsburgh compound B and fluorodeoxyglucose positron emission tomography in relation to education in mild Alzheimer's disease. Annals of Neurology, 63, 112–118. doi: 10.1002/ana.21212 Kendall, J., & Hatton, D. (2002). Racism as a source of health disparity in families with children with attention deficit hyperactivity disorder. Advances in Nursing Science, 25, 22– 39. Kertzner, R. M., Goetz, R., Todak, G., & Cooper, T. (1993). Cortisol levels, immune status, and mood in homosexual men with and without HIV infection. American Journal of Psychiatry, 150, 1674–1678. Kiecolt-Glaser, J. K., Gouin, J. P., Weng, N. P., Malarkey, W. B., Beversdorf, D. Q., & Glaser, R. (2011). Childhood adversity heightens the impact of later-life caregiving stress on telomere length and inflammation. Psychosomatic Medicine, 73, 16–22. doi: 10.1097/PSY.0b013e31820573b6 Kim, H. S. (2008). Culture and the cognitive and neuroendocrine responses to speech. Journal of Personality and Social Psychology, 94, 32–47. doi: 10.1037/0022–3514.94.1.32 Kinnally, E. L., Capitanio, J. P., Leibel, R., Deng, L., Leduc, C., Haghighi, F., & Mann, J. J. (2010). Epigenetic regulation of serotonin transporter expression and behavior in infant rhesus macaques. Genes, Brain and Behavior, 9, 575–582. doi: 10.1111/j.1601–183X.2010.00588.x Kirmayer, L., & Paul, K. (2007). Mental health, social support and community wellness. Nunavik Inuit Health Survey 2004. Kirschbaum, C., Tietze, A., Skoluda, N., & Dettenborn, L. (2009). Hair as a retrospective calendar of cortisol production—Increased cortisol incorporation into hair in the third trimester of pregnancy. Psychoneuroendocrinology, 34, 32–37. doi: 10.1016/j.psyneuen.2008.08.024 Kiuchi, T., Lee, H., & Mikami, T. (2012). Regular exercise cures depression-like behavior via

VEGF-Flk-1 signaling in chronically stressed mice. Neuroscience, 207, 208–217. doi: 10.1016/j.neuroscience.2012.01.023 Knoth, R., Singec, I., Ditter, M., Pantazis, G., Capetian, P., Meyer, R. P.,…& Kempermann, G. (2010). Murine features of neurogenesis in the human hippocampus across the lifespan from 0 to 100 years. PLoS One, 5, e8809. doi: 10.1371/journal.pone.0008809 Kogure, D., Matsuda, H., Ohnishi, T., Asada, T., Uno, M., Kunihiro, T.,…& Takasaki, M. (2000). Longitudinal evaluation of early Alzheimer's disease using brain perfusion SPECT. Journal of Nuclear Medicine, 41, 1155–1162. Kondo, K., Niino, M., & Shido, K. (1994). A case-control study of Alzheimer's disease in Japan—significance of life-styles. Dementia, 5, 314–326. doi: 10.1159/000106741 Koopman, W. J., Willems, P. H., & Smeitink, J. A. (2012). Monogenic mitochondrial disorders. New England Journal of Medicine, 366, 1132–1141. doi: 10.1056/NEJMra1012478 Korte, S. M., Koolhaas, J. M., Wingfield, J. C., & McEwen, B. S. (2005). The Darwinian concept of stress: Benefits of allostasis and costs of allostatic load and the trade-offs in health and disease. Neuroscience and Biobehavioral Reviews, 29, 3–38. doi: 10.1016/j.neubiorev.2004.08.009 Kramer, A. F., & Erickson, K. I. (2007). Capitalizing on cortical plasticity: Influence of physical activity on cognition and brain function. Trends in Cognitive Science, 11, 342–348. doi: 10.1016/j.tics.2007.06.009 Kramer, A. F., Hahn, S., Cohen, N. J., Banich, M. T., McAuley, E., Harrison, C. R.,… Colcombe, A. (1999). Ageing, fitness and neurocognitive function. Nature, 400, 418–419. doi: 10.1038/22682 Kroger, E., Andel, R., Lindsay, J., Benounissa, Z., Verreault, R., & Laurin, D. (2008). Is complexity of work associated with risk of dementia? The Canadian Study of Health and Aging. American Journal of Epidemiology, 167, 820–830. doi: 10.1093/aje/kwm382 Kronenberg, G., Bick-Sander, A., Bunk, E., Wolf, C., Ehninger, D., & Kempermann, G. (2006). Physical exercise prevents age-related decline in precursor cell activity in the mouse dentate gyrus. Neurobiology of Aging, 27, 1505–1513. doi: 10.1016/j.neurobiolaging.2005.09.016 Kubzansky, L. D., Kawachi, I., & Sparrow, D. (1999). Socioeconomic status, hostility, and risk factor clustering in the Normative Aging Study: Any help from the concept of allostatic load? Annals of Behavioral Medicine, 21, 330–338. Kuhn, H. G., Dickinson-Anson, H., & Gage, F. H. (1996). Neurogenesis in the dentate gyrus of the adult rat: age-related decrease of neuronal progenitor proliferation. Journal of Neuroscience, 16, 2027–2033.

Kyrou, I., & Tsigos, C. (2007). Stress mechanisms and metabolic complications. Hormones and Metabolic Research, 39, 430–438. doi: 10.1055/s-2007–981462 Kyrou, I., & Tsigos, C. (2009). Stress hormones: Physiological stress and regulation of metabolism. Current Opinion in Pharmacology, 9, 787–793. doi: 10.1016/j.coph.2009.08.007 Labonte, B., Suderman, M., Maussion, G., Navaro, L., Yerko, V., Mahar, I.,…Turecki, G. (2012). Genome-wide epigenetic regulation by early-life trauma. Archives of General Psychiatry, 69, 722–731. doi: 10.1001/archgenpsychiatry.2011.2287 Lane, N., & Martin, W. (2010). The energetics of genome complexity. Nature, 467, 929–934. doi: 10.1038/nature09486 Lanza, I. R., Short, D. K., Short, K. R., Raghavakaimal, S., Basu, R., Joyner, M. J.,…Nair, K. S. (2008). Endurance exercise as a countermeasure for aging. Diabetes, 57, 2933–2942. doi: 10.2337/db08–0349 Larson, E. B., Wang, L., Bowen, J. D., McCormick, W. C., Teri, L., Crane, P., & Kukull, W. (2006). Exercise is associated with reduced risk for incident dementia among persons 65 years of age and older. Annals of Internal Medicine, 144, 73–81. doi: 10.7326/0003–4819–144–2– 200601170–00004 Launer, L. J., Andersen, K., Dewey, M. E., Letenneur, L., Ott, A., Amaducci, L. A.,…Hofman, A. (1999). Rates and risk factors for dementia and Alzheimer's disease: Results from EURODEM pooled analyses. EURODEM Incidence Research Group and Work Groups. European Studies of Dementia. Neurology, 52, 78–84. doi: 10.1212/WNL.52.1.78 Lee, J., Duan, W., Long, J. M., Ingram, D. K., & Mattson, M. P. (2000). Dietary restriction increases the number of newly generated neural cells, and induces BDNF expression, in the dentate gyrus of rats. Journal of Molecular Neuroscience, 15, 99–108. doi: 10.1385/JMN:15:2:99 Lee, J., Duan, W., & Mattson, M. P. (2002). Evidence that brain-derived neurotrophic factor is required for basal neurogenesis and mediates, in part, the enhancement of neurogenesis by dietary restriction in the hippocampus of adult mice. Journal of Neurochemistry, 82, 1367– 1375. doi: 10.1046/j.1471–4159.2002.01085.x Lee, S. R., Kim, H. K., Song, I. S., Youm, J., Dizon, L. A., Jeong, S. H.,…Han, J. (2013). Glucocorticoids and their receptors: Insights into specific roles in mitochondria. Progress in Biophysiological and Molecular Biology, 112, 44–54. doi: 10.1016/j.pbiomolbio.2013.04.001 Legato, M. J. (2010). The allostatic load: How stress makes us sick. Gender Medicine, 7, 458–460. doi: 10.1016/j.genm.2010.10.001 Leserman, J., Petitto, J. M., Golden, R. N., Gaynes, B. N., Gu, H., Perkins, D. O.,…Evans, D.

L. (2000). Impact of stressful life events, depression, social support, coping, and cortisol on progression to AIDS. American Journal of Psychiatry, 157, 1221–1228. Lewis, T. T., Everson-Rose, S. A., Powell, L. H., Matthews, K. A., Brown, C., Karavolos, K., …Wesley, D. (2006). Chronic exposure to everyday discrimination and coronary artery calcification in African-American women: The SWAN Heart Study. Psychosomatic Medicine, 68, 362–368. doi: 10.1097/01.psy.0000221360.94700.16 Li, L. A., & Wang, P. W. (2005). PCB126 induces differential changes in androgen, cortisol, and aldosterone biosynthesis in human adrenocortical H295R cells. Toxicological Sciences, 85, 530–540. Li, Y., Luikart, B. W., Birnbaum, S., Chen, J., Kwon, C. H., Kernie, S. G.,…Parada, L. F. (2008). TrkB regulates hippocampal neurogenesis and governs sensitivity to antidepressive treatment. Neuron, 59, 399–412. doi: 10.1016/j.neuron.2008.06.023 Licinio, J., Dong, C., & Wong, M. L. (2009). Novel sequence variations in the brain-derived neurotrophic factor gene and association with major depression and antidepressant treatment response. Archives of General Psychiatry, 66, 488–497. doi: 10.1001/archgenpsychiatry.2009.38 Liesa, M., & Shirihai, O. S. (2013). Mitochondrial dynamics in the regulation of nutrient utilization and energy expenditure. Cell Metabolism, 17, 491–506. doi: 10.1016/j.cmet.2013.03.002 Lindqvist, A., Mohapel, P., Bouter, B., Frielingsdorf, H., Pizzo, D., Brundin, P., & ErlansonAlbertsson, C. (2006). High-fat diet impairs hippocampal neurogenesis in male rats. European Journal of Neurology, 13, 1385–1388. doi: 10.1111/j.1468–1331.2006.01500.x Liou, C. W., Lin, T. K., Huei Weng, H., Lee, C. F., Chen, T. L., Wei, Y. H.,…Wang, P. W. (2007). A common mitochondrial DNA variant and increased body mass index as associated factors for development of type 2 diabetes: Additive effects of genetic and environmental factors. Journal of Clinical Endocrinology and Metabolism, 92, 235–239. doi: 10.1210/jc.2006–0653 Lipowicz, A., Szklarska, A., & Malina, R. M. (2013). Allostatic load and socioeconomic status in Polish adult men. Journal of Biosocial Science, 46, 1–13. doi: 10.1017/S0021932013000345 Little, J. P., Safdar, A., Benton, C. R., & Wright, D. C. (2011). Skeletal muscle and beyond: the role of exercise as a mediator of systemic mitochondrial biogenesis. Applied Physiology, Nutrition, and Metabolism, 36, 598–607. doi: 10.1139/h11–076 Liu, Z., & Butow, R. A. (2006). Mitochondrial retrograde signaling. Annual Review of Genetics, 40, 159–185. doi: 10.1146/annurev.genet.40.110405.090613 Loucks, E. B., Juster, R. P., & Pruessner, J. (2008). Neuroendocrine biomarkers, allostatic

load, and the challenge of measurement: A commentary on Gersten. Social Science and Medicine, 66, 525–530. doi: 10.1016/j.socscimed.2007.09.006 Lovallo, W. R., Farag, N. H., Sorocco, K. H., Cohoon, A. J., & Vincent, A. S. (2012). Lifetime adversity leads to blunted stress axis reactivity: Studies from the Oklahoma Family Health Patterns Project. Biological Psychiatry, 71, 344–349. doi: 10.1016/j.biopsych.2011.10.018 Love, O. P., Shutt, L. J., Silfies, J. S., Bortolotti, G. R., Smits, J. E. G., & Bird, D. M. (2003). Effects of dietary PCB exposure on adrenocortical function in captive American kestrels (Falco sparverius). Ecotoxicology, 12, 199–208. Lundberg, U. (2005). Stress hormones in health and illness: The roles of work and gender. Psychoneuroendocrinology, 30, 1017–1021. doi: 10.1016/j.psyneuen.2005.03.014 Lundberg, U., & Frankenhaeuser, M. (1999). Stress and workload of men and women in highranking positions. Journal of Occupational Health Psychology, 4, 142–151. doi: 10.1037/1076–8998.4.2.142 Lupien, S. J., de Leon, M., de Santi, S., Convit, A., Tarshish, C., Nair, N. P.,…Meaney, M. J. (1998). Cortisol levels during human aging predict hippocampal atrophy and memory deficits. Nature Neuroscience, 1, 69–73. doi: 10.1038/271 Lupien, S. J., Evans, A., Lord, C., Miles, J., Pruessner, M., Pike, B., & Pruessner, J. C. (2007). Hippocampal volume is as variable in young as in older adults: implications for the notion of hippocampal atrophy in humans. Neuroimage, 34, 479–485. doi: 10.1016/j.neuroimage.2006.09.041 Lupien, S. J., King, S., Meaney, M. J., & McEwen, B. S. (2001). Can poverty get under your skin? Basal cortisol levels and cognitive function in children from low and high socioeconomic status. Development and Psychopathology, 13, 653–676. doi: 10.1017/S0954579401003133 Lupien, S. J., McEwen, B. S., Gunnar, M. R., & Heim, C. (2009). Effects of stress throughout the lifespan on the brain, behaviour and cognition. Nature Reviews Neuroscience, 10, 434– 445. doi: 10.1038/nrn2639 Lupien, S. J., Ouellet-Morin, I., Hupbach, A., Walker, D., Tu, M. T., Buss, C.,…McEwen, B. (2006). Beyond the stress concept: Allostatic load—a developmental biological and cognitive perspective. In D. Cicchetti (Ed.) Handbook series on developmental psychopathology (pp. 784–809). Hoboken, NJ, US: John Wiley & Sons Inc. Luthar, S. S., Cicchetti, D., & Becker, B. (2000). The construct of resilience: a critical evaluation and guidelines for future work. Child Development, 71, 543–562. doi: 10.1111/1467–8624.00164 Maggio, N., & Segal, M. (2007). Striking variations in corticosteroid modulation of long-term potentiation along the septotemporal axis of the hippocampus. Journal of Neuroscience, 27,

5757–5765. doi: 10.1523/JNEUROSCI.0155–07.2007 Maguire, E. A., Gadian, D. G., Johnsrude, I. S., Good, C. D., Ashburner, J., Frackowiak, R. S., & Frith, C. D. (2000). Navigation-related structural change in the hippocampi of taxi drivers. Proceedings of the National Academy of Sciences of the USA, 97, 4398–4403. doi: 10.1073/pnas.070039597 Mahar, I., Tan, S., Davoli, M. A., Dominguez-Lopez, S., Qiang, C., Rachalski, A.,… Mechawar, N. (2011). Subchronic peripheral neuregulin-1 increases ventral hippocampal neurogenesis and induces antidepressant-like effects. PLoS One, 6, e26610. doi: 10.1371/journal.pone.0026610 Malberg, J. E., Eisch, A. J., Nestler, E. J., & Duman, R. S. (2000). Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. Journal of Neuroscience, 20, 9104–9110. Manoli, I., Alesci, S., Blackman, M. R., Su, Y. A., Rennert, O. M., & Chrousos, G. P. (2007). Mitochondria as key components of the stress response. Trends in Endocrinology and Metabolism, 18, 190–198. doi: 10.1016/j.tem.2007.04.004 Marmot, M. (2005). Social determinants of health inequalities. Lancet, 365, 1099–1104. doi: 10.1016/S0140–6736(05)71146–6 Marmot, M. (2011). Social determinants and the health of Indigenous Australians. Medical Journal of Australia, 194, 512–513. Mason, J. W. (1968). A review of psychoendocrine research on the sympathetic-adrenal medullary system. Psychosomatic Medicine, 30, Suppl: 631–653. Masten, A. S., Best, K., & Garmezy, N. (1990). Resilience and development: Contributions from the study of children who overcome adversity. Development and Psychopathology, 2, 425–444. doi: 10.1017/S0954579400005812 Mattei, J., Bhupathiraju, S., & Tucker, K. L. (2013). Higher adherence to a diet score based on American Heart Association recommendations is associated with lower odds of allostatic load and metabolic syndrome in Puerto Rican adults. Journal of Nutrition, 143, 1753–1759. doi: 10.3945/jn.113.180141 Mattei, J., Noel, S. E., & Tucker, K. L. (2011). A meat, processed meat, and French fries dietary pattern is associated with high allostatic load in Puerto Rican older adults. Journal of the American Diabetes Association, 111, 1498–1506. doi: 10.1016/j.jada.2011.07.006 Mattson, M. P. (2012). Energy intake and exercise as determinants of brain health and vulnerability to injury and disease. Cell Metabolism, 16, 706–722. doi: 10.1016/j.cmet.2012.08.012 Mattson, M. P., Chan, S. L., & Duan, W. (2002). Modification of brain aging and

neurodegenerative disorders by genes, diet, and behavior. Physiological Reviews, 82, 637– 672. doi: 10.1152/physrev.00004.2002 Mays, V. M., & Cochran, S. D. (2001). Mental health correlates of perceived discrimination among lesbian, gay, and bisexual adults in the United States. American Journal of Public Health, 91, 1869–1876. Mays, V. M., Cochran, S. D., & Barnes, N. W. (2007). Race, race-based discrimination, and health outcomes among African Americans. Annual Review of Psychology, 58, 201–225. doi: 10.1146/annurev.psych.57.102904.190212 McEwen, B. S. (1998a). Protective and damaging effects of stress mediators. New England Journal of Medicine, 338, 171–179. doi: 10.1056/NEJM199801153380307 McEwen, B. S. (1998b). Stress, adaptation, and disease. Allostasis and allostatic load. Annals of the New York Academy of Science, 840, 33–44. doi: 10.1111/j.1749– 6632.1998.tb09546.x McEwen, B. S. (2001). From molecules to mind. Stress, individual differences, and the social environment. Annals of the New York Academy of Science, 935, 42–49. doi: 10.1111/j.1749– 6632.2001.tb03469.x McEwen, B. S. (2003). Interacting mediators of allostasis and allostatic load: toward an understanding of resilience in aging. Metabolism, 52, 10–16. doi: 10.1053/S0026– 0495(03)00295–6 McEwen, B. S. (2006). Sleep deprivation as a neurobiologic and physiologic stressor: Allostasis and allostatic load. Metabolism, 55, S20–S23. doi: 10.1016/j.metabol.2006.07.008 McEwen, B. S. (2007). Physiology and neurobiology of stress and adaptation: Central role of the brain. Physiological Reviews, 87, 873–904. doi: 10.1152/physrev.00041.2006 McEwen, B. S. (2009). The brain is the central organ of stress and adaptation. Neuroimage, 47, 911–913. doi: 10.1016/j.neuroimage.2009.05.071 McEwen, B. S. (2012). Brain on stress: How the social environment gets under the skin. Proceedings of the National Academy of Sciences of the USA, 109, Suppl. 2, 17180–17185. doi: 10.1073/pnas.1121254109 McEwen, B. S., & Gianaros, P. J. (2011). Stress- and allostasis-induced brain plasticity. Annual Review of Medicine, 62, 431–445. doi: 10.1146/annurev-med-052209–100430 McEwen, B. S., & Seeman, T. (1999). Protective and damaging effects of mediators of stress. Elaborating and testing the concepts of allostasis and allostatic load. Annals of the New York Academy of Science, 896, 30–47. doi: 10.1111/j.1749–6632.1999.tb08103.x McEwen, B. S., & Stellar, E. (1993). Stress and the individual. Mechanisms leading to disease. Archives of Internal Medicine, 153, 2093–2101. doi:

10.1001/archinte.1993.00410180039004 McEwen, B. S., & Tucker, P. (2011). Critical biological pathways for chronic psychosocial stress and research oppurtunities to advance the consideration of stress in chemical risk assessment. American Journal of Public Health EPub, S131–S1319. doi: 10.2105/AJPH.2011.300270 McEwen, B. S., & Wingfield, J. C. (2003). The concept of allostasis in biology and biomedicine. Hormones and Behavior, 43, 2–15. doi: 10.1016/S0018–506X(02)00024–7 McGhee, R. (1978). Canadian Arctic prehistory. New York, NY: Van Nostrand Reinhold. Meaney, M. J. (2001). Maternal care, gene expression, and the transmission of individual differences in stress reactivity across generations. Annual Review of Neuroscience, 24, 1161– 1192. doi: 10.1146/annurev.neuro.24.1.1161 Meaney, M. J., & Szyf, M. (2005). Environmental programming of stress responses through DNA methylation: Life at the interface between a dynamic environment and a fixed genome. Dialogues in Clinical Neuroscience, 7, 103–123. Meerlo, P., Mistlberger, R. E., Jacobs, B. L., Heller, H. C., & McGinty, D. (2009). New neurons in the adult brain: The role of sleep and consequences of sleep loss. Sleep Medicine Reviews, 13, 187–194. doi: 10.1016/j.smrv.2008.07.004 Merkin, S. S., Basurto-Davila, R., Karlamangla, A., Bird, C. E., Lurie, N., Escarce, J., & Seeman, T. (2009). Neighborhoods and cumulative biological risk profiles by race/ethnicity in a national sample of U.S. adults: NHANES III. Annals of Epidemiology, 19, 194–201. doi: 10.1016/j.annepidem.2008.12.006 Meserve, L. A., Murray, B. A., & Landis, J. A. (1992). Influence of maternal ingestion of Aroclor 1254 (PCB) or FireMaster BP-6 (PBB) on unstimulated and stimulated corticosterone levels in young rats. Bulletin of Environmental Contamination and Toxicology, 48, 715–720. Meyer, I. H. (1995). Minority stress and mental health in gay men. Journal of Health and Social Behavior, 36, 38–56. Meyer, I. H. (2003). Prejudice, social stress, and mental health in lesbian, gay, and bisexual populations: Conceptual issues and research evidence. Psychological Bulletin, 129, 674–697. Meyer, I. H. (2007). Prejudice and discrimination as social stressors. In I. H. Meyer & M. E. Northridge (Eds.), The health of sexual minorities: Public health perspectives on lesbian, gay, bisexual, and transgender populations (pp. 242–267). New York, NY: Springer. Miller, G. E., & Chen, E. (2010). Harsh family climate in early life presages the emergence of a proinflammatory phenotype in adolescence. Psychological Science, 21, 848–856. doi: 10.1177/0956797610370161 Miller, G. E., Chen, E., & Zhou, E. S. (2007). If it goes up, must it come down? Chronic stress

and the hypothalamic-pituitary-adrenocortical axis in humans. Psychosomatic Medicine, 133, 25–45. doi: 10.1037/0033–2909.133.1.25 Mineur, Y. S., Belzung, C., & Crusio, W. E. (2007). Functional implications of decreases in neurogenesis following chronic mild stress in mice. Neuroscience, 150, 251–259. doi: 10.1016/j.neuroscience.2007.09.045 Mirescu, C., Peters, J. D., Noiman, L., & Gould, E. (2006). Sleep deprivation inhibits adult neurogenesis in the hippocampus by elevating glucocorticoids. Proceedings of the National Academy of Sciences of the USA, 103, 19170–19175. doi: 10.1073/pnas.0608644103 Morava, E., & Kozicz, T. (2013). Mitochondria and the economy of stress (mal)adaptation. Neuroscience and Biobehavioral Reviews, 37, 668–680. doi: 10.1016/j.neubiorev.2013.02.005 Morris, J. F., Waldo, C. R., & Rothblum, E. D. (2001). A model of predictors and outcomes of outness among lesbian and bisexual women. American Journal of Orthopsychiatry, 71, 61– 71. Mortel, K. F., Meyer, J. S., Herod, B., & Thornby, J. (1995). Education and occupation as risk factors for dementias of the Alzheimer and ischemic vascular types. Dementia, 6, 55–62. doi: 10.1159/000106922 Muckle, G., Ayotte, P., Dewailly, E. E., Jacobson, S. W., & Jacobson, J. L. (2001). Prenatal exposure of the northern Quebec Inuit infants to environmental contaminants. Environmental Health Perspectives, 109, 1291–1299. Mueller, A. D., Mear, R. J., & Mistlberger, R. E. (2011). Inhibition of hippocampal neurogenesis by sleep deprivation is independent of circadian disruption and melatonin suppression. Neuroscience, 193, 170–181. doi: 10.1016/j.neuroscience.2011.07.019 Mueller, A. D., Pollock, M. S., Lieblich, S. E., Epp, J. R., Galea, L. A., & Mistlberger, R. E. (2008). Sleep deprivation can inhibit adult hippocampal neurogenesis independent of adrenal stress hormones. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology, 294, R1693–R1703. doi: 10.1152/ajpregu.00858.2007 Murgatroyd, C., Patchev, A. V., Wu, Y., Micale, V., Bockmuhl, Y., Fischer, D.,…Spengler, D. (2009). Dynamic DNA methylation programs persistent adverse effects of early-life stress. Nature Neuroscience, 12, 1559–1566. doi: 10.1038/nn.2436 Musazzi, L., Cattaneo, A., Tardito, D., Barbon, A., Gennarelli, M., Barlati, S.,…Popoli, M. (2009). Early raise of BDNF in hippocampus suggests induction of posttranscriptional mechanisms by antidepressants. BMC Neuroscience, 10–48, 48. doi: 10.1186/1471–2202–10– 48 Mustanski, B., Garofalo, R., Herrick, A., & Donenberg, G. (2007). Psychosocial health problems increase risk for HIV among urban young men who have sex with men: Preliminary

evidence of a syndemic in need of attention. Annals of Behavioral Medicine, 34, 37–45. Myers, H. F. (2009). Ethnicity- and socio-economic status-related stresses in context: An integrative review and conceptual model. Journal of Behavioral Medicine, 32, 9–19. doi: 10.1007/s10865–008–9181–4 Nasca, C., Xenos, D., Barone, Y., Caruso, A., Scaccianoce, S., Matrisciano, F.,…Nicoletti, F. (2013). L-acetylcarnitine causes rapid antidepressant effects through the epigenetic induction of mGlu2 receptors. Proceedings of the National Academy of Sciences of the USA, 110, 4804–4809. doi: 10.1073/pnas.1216100110 Naumova, O. Y., Lee, M., Koposov, R., Szyf, M., Dozier, M., & Grigorenko, E. L. (2011). Differential patterns of whole-genome DNA methylation in institutionalized children and children raised by their biological parents. Development and Psychopathology, 24, 143–155. doi: 10.1017/S0954579411000605 Needham, B. L., Adler, N., Gregorich, S., Rehkopf, D., Lin, J., Blackburn, E. H., & Epel, E. S. (2013). Socioeconomic status, health behavior, and leukocyte telomere length in the National Health and Nutrition Examination Survey, 1999–2002. Social Science and Medicine, 85, 1–8. doi: 10.1016/j.socscimed.2013.02.023 Needham, B. L., Fernandez, J. R., Lin, J., Epel, E. S., & Blackburn, E. H. (2012). Socioeconomic status and cell aging in children. Social Science and Medicine, 74, 1948– 1951. doi: 10.1016/j.socscimed.2012.02.019 Neigh, G. N., Gillespie, C. F., & Nemeroff, C. B. (2009). The neurobiological toll of child abuse and neglect. Trauma, Violence, and Abuse, 10, 389–410. doi: 10.1177/1524838009339758 Nelson, K. M., Reiber, G., Kohler, T., & Boyko, E. J. (2007). Peripheral arterial disease in a multiethnic national sample: the role of conventional risk factors and allostatic load. Ethnicity and Disease, 17, 669–675. Nettles, K. W., Pesold, C., & Goldman, M. B. (2000). Influence of the ventral hippocampal formation on plasma vasopressin, hypothalamic-pituitary-adrenal axis, and behavioral responses to novel acoustic stress. Brain Research, 858, 181–190. doi: 10.1016/S0006– 8993(99)02281–7 Ngandu, T., von Strauss, E., Helkala, E. L., Winblad, B., Nissinen, A., Tuomilehto, J.,… Kivipelto, M. (2007). Education and dementia: what lies behind the association? Neurology, 69, 1442–1450. doi: 10.1212/01.wnl.0000277456.29440.16 Nibuya, M., Morinobu, S., & Duman, R. S. (1995). Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. Journal of Neuroscience, 15, 7539–7547. Niemi, A. K., & Majamaa, K. (2005). Mitochondrial DNA and ACTN3 genotypes in Finnish

elite endurance and sprint athletes. European Journal of Human Genetics, 13, 965–969. doi: 10.1038/sj.ejhg.5201438 Nyberg, C. H. (2012). Diurnal cortisol rhythms in Tsimane' Amazonian foragers: new insights into ecological HPA axis research. Psychoneuroendocrinology, 37, 178–190. doi: 10.1016/j.psyneuen.2011.06.002 O'Keefe, J., & Nadel, L. (1978). The hippocampus as a cognitive map. Oxford, UK: Clarendon Press. Oskam, I. C., Ropstad, E., Lie, E., Derocher, A. E., Wiig, O., Dahl, E.,…Skaare, J. U. (2004). Organochlorines affect the steroid hormone cortisol in free-ranging polar bears (Ursus maritimus) at Svalbard, Norway. Journal of Toxicology and Environmental Health, Part A, Current Issues, 67, 959–977. Ott, A., Breteler, M. M., van Harskamp, F., Claus, J. J., van der Cammen, T. J., Grobbee, D. E., & Hofman, A. (1995). Prevalence of Alzheimer's disease and vascular dementia: Association with education. The Rotterdam study. British Medical Journal, 310, 970–973. doi: 10.1136/bmj.310.6985.970 Ouellet-Morin, I., Danese A., Bowes L., Shakoor S., Ambler A., Pariante C.,…Arseneault L. (2011). A discordant MZ twin design shows blunted cortisol reactivity among bullied children. Journal of the American Academy of Child & Adolescent Psychiatry, 50, 574–582. doi: 10.1016/j.jaac.2011.02.015 Ouellet-Morin, I., Wong, C., Danese, A., Caspi, A., Moffitt, T. E., Mill, J., & Arseneault, L. (2013). DNA methylation in candidate genes for HPA axis: a discordant MZ twin design. Psychological Medicine, 43, 1813–1823. doi: 10.1017/S0033291712002784 Paizanis, E., Renoir, T., Lelievre, V., Saurini, F., Melfort, M., Gabriel, C.,…Lanfumey, L. (2010). Behavioural and neuroplastic effects of the new-generation antidepressant agomelatine compared to fluoxetine in glucocorticoid receptor-impaired mice. International Journal of Neuropsychopharmacology, 13, 759–774. doi: 10.1017/S1461145709990514 Parente, V., Hale, L., & Palermo, T. (2013). Association between breast cancer and allostatic load by race: National Health and Nutrition Examination Survey 1999–2008. PsychoOncology, 22, 621–628. doi: 10.1002/pon.3044 Peng, S., Wuu, J., Mufson, E. J., & Fahnestock, M. (2005). Precursor form of brain-derived neurotrophic factor and mature brain-derived neurotrophic factor are decreased in the preclinical stages of Alzheimer's disease. Journal of Neurochemistry, 93, 1412–1421. doi: 10.1111/j.1471–4159.2005.03135.x Perera, T. D., Dwork, A. J., Keegan, K. A., Thirumangalakudi, L., Lipira, C. M., Joyce, N.,… Coplan, J. D. (2011). Necessity of hippocampal neurogenesis for the therapeutic action of antidepressants in adult nonhuman primates. PLoS One, 6, e17600. doi:

10.1371/journal.pone.0017600 Perfilieva, E., Risedal, A., Nyberg, J., Johansson, B. B., & Eriksson, P. S. (2001). Gender and strain influence on neurogenesis in dentate gyrus of young rats. Journal of Cerebral Blood Flow and Metabolism, 21, 211–217. doi: 10.1097/00004647–200103000–00004 Perlis, R. H., Ruderfer, D., Hamilton, S. P., & Ernst, C. (2012). Copy number variation in subjects with major depressive disorder who attempted suicide. PLoS One, 7, e46315. doi: 10.1371/journal.pone.0046315 Perneczky, R., Drzezga, A., Diehl-Schmid, J., Schmid, G., Wohlschlager, A., Kars, S.,…Kurz, A. (2006). Schooling mediates brain reserve in Alzheimer's disease: Findings of fluoro-deoxyglucose-positron emission tomography. Journal of Neurology, Neurosurgery, and Psychiatry, 77, 1060–1063. doi: 10.1136/jnnp.2006.094714 Petersen, K. F., Dufour, S., Befroy, D., Garcia, R., & Shulman, G. I. (2004). Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes. New England Journal of Medicine, 350, 664–671. doi: 10.1056/NEJMoa031314 Peterson, J. L., & Jones, K. T. (2009). HIV prevention for black men who have sex with men in the United States. American Journal of Public Health, 99, 976–980. doi: 10.2105/AJPH.2008.143214 Picard, M. (2011). Pathways to aging: the mitochondrion at the intersection of biological and psychosocial sciences. Journal of Aging Research 2011, 814096. doi: 10.4061/2011/814096 Picard, M., Juster, R. P., & McEwen, B. S. (2014). Mitochondrial allostatic load puts the 'gluc' back in glucocorticoids. Nature Reviews Endocrinology, 10, 303–310. doi: 10.1038/nrendo.2014.22 Picard, M., & McEwen, B. S. (2014). Mitochondria impact brain function and cognition. Proceedings from the National Academy of Sciences of the USA, 111, 7–8. doi: 10.1073/pnas.1321881111 Picard, M., & Turnbull, D. M. (2013). Linking the metabolic state and mitochondrial DNA in chronic disease, health and aging. Diabetes, 62, 672–678. doi: 10.2337/db12–1203 Pijl, H. (2012). Longevity. The allostatic load of dietary restriction. Physiology and Behavior, 106, 51–57. doi: 10.1016/j.physbeh.2011.05.030 Pike, I. L., & Williams, S. R. (2006). Incorporating psychosocial health into biocultural models: preliminary findings from Turkana women of Kenya. American Journal of Human Biology, 18, 729–740. doi: 10.1002/ajhb.20548 Pilote, L., Dasgupta, K., Guru, V., Humphries, K. H., McGrath, J., Norris, C.,…Tagalakis, V. (2007). A comprehensive view of sex-specific issues related to cardiovascular disease. Canadian Medical Association Journal, 176, S1–S44. doi: 10.1503/cmaj.051455

Plusquellec, P., Muckle, G., Dewailly, E., Ayotte, P., Begin, G., Desrosiers, C.,…Poitras, K. (2010). The relation of environmental contaminants exposure to behavioral indicators in Inuit preschoolers in Arctic Quebec. Neurotoxicology, 31, 17–25. doi: 10.1016/j.neuro.2009.10.008 Plusquellec, P., Muckle, G., Dewailly, E., Ayotte, P., Jacobson, S. W., & Jacobson, J. L. (2007). The relation of low-level prenatal lead exposure to behavioral indicators of attention in Inuit infants in Arctic Quebec. Neurotoxicology and Teratology, 29, 527–537. doi: 10.1016/j.ntt.2007.07.002 Potter, G. G., Helms, M. J., Burke, J. R., Steffens, D. C., & Plassman, B. L. (2007). Job demands and dementia risk among male twin pairs. Alzheimer's and Dementia, 3, 192–199. doi: 10.1016/j.jalz.2007.04.377 Power, C., Thomas, C., Li, L., & Hertzman, C. (2012). Childhood psychosocial adversity and adult cortisol patterns. British Journal of Psychiatry, 201, 199–206. doi: 10.1192/bjp.bp.111.096032 Provencal, N., Suderman, M. J., Guillemin, C., Massart, R., Ruggiero, A., Wang, D.,…Szyf, M. (2012). The signature of maternal rearing in the methylome in rhesus macaque prefrontal cortex and T cells. Journal of Neuroscience, 32, 15626–15642. doi: 10.1523/JNEUROSCI.1470– 12.2012 Pruessner, J. C., Baldwin, M. W., Dedovic, K., Renwick, R., Mahani, N. K., Lord, C.,… Lupien, S. (2005). Self-esteem, locus of control, hippocampal volume, and cortisol regulation in young and old adulthood. Neuroimage, 28, 815–826. doi: 10.1016/j.neuroimage.2005.06.014 Pruessner, J. C., Champagne, F., Meaney, M. J., & Dagher, A. (2004). Dopamine release in response to a psychological stress in humans and its relationship to early life maternal care: a positron emission tomography study using [11C]raclopride. Journal of Neuroscience, 24, 2825–2831. doi: 10.1523/JNEUROSCI.3422–03.2004 Pruessner, J. C., Dedovic, K., Khalili-Mahani, N., Engert, V., Pruessner, M., Buss, C.,… Lupien, S. (2008). Deactivation of the limbic system during acute psychosocial stress: Evidence from positron emission tomography and functional magnetic resonance imaging studies. Biological Psychiatry, 63, 234–240. doi: 10.1016/j.biopsych.2007.04.041 Pruessner, J. C., Dedovic, K., Pruessner, M., Lord, C., Buss, C., Collins, L.,…Lupien, S. J. (2010). Stress regulation in the central nervous system: evidence from structural and functional neuroimaging studies in human populations—2008 Curt Richter Award Winner. Psychoneuroendocrinology. 35, 179–191. doi: 10.1016/j.psyneuen.2009.02.016 Pruessner, J. C., Hellhammer, D. H., & Kirschbaum, C. (1999). Low self-esteem, induced failure and the adrenocortical stress response. Personality and Individual Differences, 27, 477–489. doi: 10.1016/S0191–8869(98)00256–6

Psarra, A. M., & Sekeris, C. E. (2009). Glucocorticoid receptors and other nuclear transcription factors in mitochondria and possible functions. Biochimica et Biophysica Acta, 1787, 431–436. doi: 10.1016/j.bbabio.2008.11.011 Puterman, E., Lin, J., Blackburn, E., O'Donovan, A., Adler, N., & Epel, E. (2010). The power of exercise: buffering the effect of chronic stress on telomere length. PLoS One, 5, e10837. doi: 10.1371/journal.pone.0010837 Qiu, C., Backman, L., Winblad, B., Aguero-Torres, H., & Fratiglioni, L. (2001). The influence of education on clinically diagnosed dementia incidence and mortality data from the Kungsholmen Project. Archives of Neurology, 58, 2034–2039. doi: 10.1001/archneur.58.12.2034 Rainisch, B. K., & Upchurch, D. M. (2013). Sociodemographic correlates of allostatic load among a national sample of adolescents: Findings from the National Health and Nutrition Examination Survey, 1999–2008. Journal of Adolescent Health, 53, 506–511. doi: 10.1016/j.jadohealth.2013.04.020 Ravussin, E., Valencia, M. E., Esparza, J., Bennett, P. H., & Schulz, L. O. (1994). Effects of a traditional lifestyle on obesity in Pima Indians. Diabetes Care, 17, 1067–1074. Reif, A., Fritzen, S., Finger, M., Strobel, A., Lauer, M., Schmitt, A., & Lesch, K. P. (2006). Neural stem cell proliferation is decreased in schizophrenia, but not in depression. Molecular Psychiatry, 11, 514–522. doi: 10.1038/sj.mp.4001791 Rettberg, J. R., Yao, J., Brinton, R. D. (2014). Estrogen: A master regulator of bioenergetic systems in the brain and body. Frontiers in Neuroendocrinology, 35, 8–30. doi: 10.1016/j.yfrne.2013.08.001 Riley, B. H. (2010). GLB adolescent's “coming out.” Journal of Child and Adolescent Psychiatric Nursing, 23, 3–10. Riva, M. A., Plusquellec, P., Juster, R. P., Laouan-Sidi, E. A., Abdous, B., Lucas, M.,… Dewailly, E. (2014) Household crowding is associated with higher allostatic load among Inuit. Journal of Epidemiology and Community Health, 68, 363–369. doi: 10.1136/jech-2013– 203270 Rochette, L., St-Laurent, D., & Plaziac, C. (2007). Socio-demographic portrait. Nunavik Inuit Health Survey 2004. Rodrigues, R. T. S., Barbosa, G. S., & Chiavone, P. A. (2013). Personality and resilience as protection against burnout in resident doctors. Revista Brasileira de Educação Médica, 37, 245–253. Roe, C. M., Mintun, M. A., D'Angelo, G., Xiong, C., Grant, E. A., & Morris, J. C. (2008). Alzheimer disease and cognitive reserve: Variation of education effect with carbon 11-labeled Pittsburgh Compound B uptake. Archives of Neurology, 65, 1467–1471. doi:

10.1001/archneur.65.11.1467 Rollins, B., Martin, M. V., Sequeira, P. A., Moon, E. A., Morgan, L. Z., Watson, S. J.,…Vawter, M. P. (2009). Mitochondrial variants in schizophrenia, bipolar disorder, and major depressive disorder. PLoS One, 4, e4913. doi: 10.1371/journal.pone.0004913 Rosal, M. C., King, J., Ma, Y., & Reed, G. W. (2004). Stress, social support, and cortisol: inverse associations? Behavioral Medicine, 30, 11–21. doi: 10.3200/BMED.30.1.11–22 Rosano, C., Venkatraman, V. K., Guralnik, J., Newman, A. B., Glynn, N. W., Launer, L.,… Aizenstein, H. (2010). Psychomotor speed and functional brain MRI 2 years after completing a physical activity treatment. Journals of Gerontology: Series A, Biological Sciences and Medical Sciences, 65, 639–647. doi: 10.1093/gerona/glq038 Ross, M. W. (1990). The relationship between life events and mental health in homosexual men. Journal of Clinical Psychology, 46, 402–411. Rossi, C., Angelucci, A., Costantin, L., Braschi, C., Mazzantini, M., Babbini, F.,…Caleo, M. (2006). Brain-derived neurotrophic factor (BDNF) is required for the enhancement of hippocampal neurogenesis following environmental enrichment. European Journal of Neuroscience, 24, 1850–1856. doi: 10.1111/j.1460–9568.2006.05059.x Roth, T. L., Lubin, F. D., Funk, A. J., & Sweatt, J. D. (2009). Lasting epigenetic influence of early-life adversity on the BDNF gene. Biological Psychiatry, 65, 760–769. doi: 10.1016/j.biopsych.2008.11.028 Roth, T. L., & Sweatt, J. D. (2011). Annual research review: Epigenetic mechanisms and environmental shaping of the brain during sensitive periods of development. Journal of Child Psychology and Psychiatry, 52, 398–408. doi: 10.1111/j.1469–7610.2010.02282.x Rowe, J. W., & Kahn, R. L. (1997). Successful aging. Gerontologist, 37, 433–440. doi: 10.1093/geront/37.4.433 Rutter, M. (1985). Resilience in the face of adversity. Protective factors and resistance to psychiatric disorder. British Journal of Psychiatry, 147, 598–611. Rutter, M. (2007). Proceeding from observed correlation to causal inference: The use of natural experiment. Perspectives in Psychological Science, 2, 377–395. doi: 10.1111/j.1745– 6916.2007.00050.x Rutter, M. (2012). Resilience as a dynamic concept. Development and Psychopathology, 24, 335–344. doi: 10.1017/S0954579412000028 Rutter, M., & Sroufe, L. A. (2000). Developmental psychopathology: Concepts and challenges. Development and Psychopathology, 12, 265–296. Sabbah, W., Watt, R. G., Sheiham, A., & Tsakos, G. (2008). Effects of allostatic load on the social gradient in ischaemic heart disease and periodontal disease: Evidence from the Third

National Health and Nutrition Examination Survey. Journal of Epidemiology and Community Health, 62, 415–420. doi: 10.1136/jech.2007.064188 Sabiston, C. M., & Crocker, P. R. (2008). Exploring self-perceptions and social influences as correlates of adolescent leisure-time physical activity. Journal of Sport and Exercise Psychology, 30, 3–22. Safren, S. A., Reisner, S. L., Herrick, A., Mimiaga, M. J., & Stall, R. D. (2010). Mental health and HIV risk in men who have sex with men. Journal of Acquired Immune Deficiency Syndromes, 55, S74–S77. Sagan, L. (1967). On the origin of mitosing cells. Journal of Theoretical Biology, 14, 255– 274. Sahin, E., & Depinho, R. A. (2010). Linking functional decline of telomeres, mitochondria and stem cells during ageing. Nature, 464, 520–528. doi: 10.1038/nature08982 Sairanen, M., Lucas, G., Ernfors, P., Castren, M., & Castren, E. (2005). Brain-derived neurotrophic factor and antidepressant drugs have different but coordinated effects on neuronal turnover, proliferation, and survival in the adult dentate gyrus. Journal of Neuroscience, 25, 1089–1094. doi: 10.1523/JNEUROSCI.3741–04.2005 Sanchez, M. M. (2006). The impact of early adverse care on HPA axis development: Nonhuman primate models. Hormones and Behavior, 50, 623–631. doi: 10.1016/j.yhbeh.2006.06.012 Santarelli, L., Saxe, M., Gross, C., Surget, A., Battaglia, F., Dulawa, S.,…Hen, R. (2003). Requirement of hippocampal neurogenesis for the behavioral effects of antidepressants. Science, 301, 805–809. doi: 10.1126/science.1083328 Sapolsky, R. M. (2000). Glucocorticoids and hippocampal atrophy in neuropsychiatric disorders. Archives of General Psychiatry, 57, 925–935. doi: 10.1001/archpsyc.57.10.925 Sapolsky, R. M. (2005). The influence of social hierarchy on primate health. Science, 308, 648–652. doi: 10.1126/science.1106477 Sapolsky, R. M., Romero, M., & Munck, A. U. (2000). How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews, 21, 55–89. doi: 10.1210/edrv.21.1.0389 Scarmeas, N., Albert, S. M., Manly, J. J., & Stern, Y. (2006). Education and rates of cognitive decline in incident Alzheimer's disease. Journal of Neurology, Neurosurgery, and Psychiatry, 77, 308–316. doi: 10.1136/jnnp.2005.072306 Scarmeas, N., Levy, G., Tang, M. X., Manly, J., & Stern, Y. (2001). Influence of leisure activity on the incidence of Alzheimer's disease. Neurology, 57, 2236–2242. doi: 10.1212/WNL.57.12.2236

Scarmeas, N., Zarahn, E., Anderson, K. E., Habeck, C. G., Hilton, J., Flynn, J.,…Stern, Y. (2003). Association of life activities with cerebral blood flow in Alzheimer disease— Implications for the cognitive reserve hypothesis. Archives of Neurology, 60, 359–365. doi: 10.1001/archneur.60.3.359 Scharfman, H., Goodman, J., Macleod, A., Phani, S., Antonelli, C., & Croll, S. (2005). Increased neurogenesis and the ectopic granule cells after intrahippocampal BDNF infusion in adult rats. Experimental Neurology, 192, 348–356. doi: 10.1016/j.expneurol.2004.11.016 Scheffler, I. E. (1999). Mitochondria. New York, NY: Wiley-Liss. Schloesser, R. J., Lehmann, M., Martinowich, K., Manji, H. K., & Herkenham, M. (2010). Environmental enrichment requires adult neurogenesis to facilitate the recovery from psychosocial stress. Molecular Psychiatry, 15, 1152–1163. doi: 10.1038/mp.2010.34 Schloesser, R. J., Manji, H. K., & Martinowich, K. (2009). Suppression of adult neurogenesis leads to an increased hypothalamo-pituitary-adrenal axis response. Neuroreport, 20, 553–557. doi: 10.1097/WNR.0b013e3283293e59 Schmidt, H. D., & Duman, R. S. (2010). Peripheral BDNF produces antidepressant-like effects in cellular and behavioral models. Neuropsychopharmacology, 35, 2378–2391. doi: 10.1038/npp.2010.114 Schmitt, L. H., Harrison, G. A., Spargo, R. M., Pollard, T., & Ungpakorn, G. (1995). Patterns of cortisol and adrenaline variation in Australian aboriginal communities of the Kimberley region. Journal of Biosocial Science, 27, 107–116. doi: 10.1017/S0021932000007045 Schoenfeld, T. J., & Gould, E. (2012). Stress, stress hormones, and adult neurogenesis. Experimental Neurology, 233, 12–21. doi: 10.1016/j.expneurol.2011.01.008 Schulz, A. J., Mentz, G., Lachance, L., Johnson, J., Gaines, C., & Israel, B. A. (2012). Associations between socioeconomic status and allostatic load: effects of neighborhood poverty and tests of mediating pathways. American Journal of Public Health, 102, 1706– 1714. doi: 10.2105/AJPH.2011.300412 Schwerzmann, K., Hoppeler, H., Kayar, S. R., & Weibel, E. R. (1989). Oxidative capacity of muscle and mitochondria: correlation of physiological, biochemical, and morphometric characteristics. Proceedings of the National Academy of Sciences of the USA, 86, 1583– 1587. Seeman, T. E., Crimmins, E., Huang, M. H., Singer, B., Bucur, A., Gruenewald, T., Berkman, L. F., & Reuben, D. B. (2004a). Cumulative biological risk and socio-economic differences in mortality: MacArthur studies of successful aging. Social Science and Medicine, 58, 1985– 1997. doi: 10.1016/S0277–9536(03)00402–7 Seeman, T. E., Glei, D., Goldman, N., Weinstein, M., Singer, B., & Lin, Y. H. (2004b). Social relationships and allostatic load in Taiwanese elderly and near elderly. Social Science and

Medicine, 59, 2245–2257. doi: 10.1016/j.socscimed.2004.03.027 Seeman, T., Gruenewald, T., Karlamangla, A., Sidney, S., Liu, K., McEwen, B., & Schwartz, J. (2010). Modeling multisystem biological risk in young adults: The Coronary Artery Risk Development in Young Adults Study. American Journal of Human Biology, 22, 463–472. doi: 10.1002/ajhb.21018 Seeman, T. E., Lusignolo, T. M., Albert, M., & Berkman, L. (2001a). Social relationships, social support, and patterns of cognitive aging in healthy, high-functioning older adults: MacArthur studies of successful aging. Health Psychology, 20, 243–255. doi: 10.1037/0278– 6133.20.4.243 Seeman, T. E., McEwen, B. S., Rowe, J. W., & Singer, B. H. (2001b). Allostatic load as a marker of cumulative biological risk: MacArthur studies of successful aging. Proceedings of the National Academy of Sciences of the USA, 98, 4770–4775. doi: 10.1073/pnas.081072698 Seeman, T., Merkin, S. S., Crimmins, E., Koretz, B., Charette, S., & Karlamangla, A. (2008). Education, income and ethnic differences in cumulative biological risk profiles in a national sample of US adults: NHANES III (1988–1994). Social Science and Medicine, 66, 72–87. doi: 10.1016/j.socscimed.2007.08.027 Seeman, E., Singer, B. H., Rowe, J., Horwitz, R. I., & McEwen, B. (1997). Price of adaptation —allostatic load and its health consequences. Archives of Internal Medicine, 157, 2259– 2268. doi: 10.1001/archinte .1997.00440400111013 Seeman, T. E., Singer, B. H., Ryff, C. D., Dienberg Love, G., & Levy-Storms, L. (2002). Social relationships, gender, and allostatic load across two age cohorts. Psychosomatic Medicine, 64, 395–406. Sequeira, A., Martin, M. V., Rollins, B., Moon, E. A., Bunney, W. E., Macciardi, F.,…Vawter, M. P. (2012). Mitochondrial mutations and polymorphisms in psychiatric disorders. Frontiers in Genetics, 3, 103. doi: 10.3389/fgene.2012.00103 Shalev, I., Moffitt, T. E., Sugden, K., Williams, B., Houts, R. M., Danese, A.,…Caspi, A. (2013). Exposure to violence during childhood is associated with telomere erosion from 5 to 10 years of age: a longitudinal study. Molecular Psychiatry, 18, 576–581. doi: 10.1038/mp.2012.32 Sharpley, M. S., Marciniak, C., Eckel-Mahan, K., McManus, M., Crimi, M., Waymire, K.,… Wallace, D. C. (2012). Heteroplasmy of mouse mtDNA is genetically unstable and results in altered behavior and cognition. Cell, 151, 333–343. doi: 10.1016/j.cell.2012.09.004 Shi, S. S., Shao, S. H., Yuan, B. P., Pan, F., & Li, Z. L. (2010). Acute stress and chronic stress change brain-derived neurotrophic factor (BDNF) and tyrosine kinase-coupled receptor (TrkB) expression in both young and aged rat hippocampus. Yonsei Medical Journal, 51, 661– 671. doi: 10.3349/ymj.2010.51.5.661

Shirayama, Y., Chen, A. C., Nakagawa, S., Russell, D. S., & Duman, R. S. (2002). Brainderived neurotrophic factor produces antidepressant effects in behavioral models of depression. Journal of Neuroscience, 22, 3251–3261. Shirtcliff, E. A., Allison, A. L., Armstrong, J. M., Slattery, M. J., Kalin, N. H., & Essex, M. J. (2011). Longitudinal stability and developmental properties of salivary cortisol levels and circadian rhythms from childhood to adolescence. Developmental Psychobiology, 493–502. doi: 10.1002/dev.20607 Shonkoff, J. P., Boyce, W. T., & McEwen, B. S. (2009). Neuroscience, molecular biology, and the childhood roots of health disparities: Building a new framework for health promotion and disease prevention. Journal of the Amercan Medical Association, 301, 2252–2259. doi: 10.1001/jama.2009.754 Short, K. R., Bigelow, M. L., Kahl, J., Singh, R., Coenen-Schimke, J., Raghavakaimal, S., & Nair, K. S. (2005). Decline in skeletal muscle mitochondrial function with aging in humans. Proceedings of the National Academy of Sciences of the USA, 102, 5618–5623. doi: 10.1073/pnas.0501559102 Silva, M. R. S., Lacharité, C., Silva, P. A., Lunardi, V. L., & Lunardi-Filho, D. (2009). Family processes that sustain resilience: A case study. Texto and Contexto Enfermagem, 18, 92–99. doi: 10.1590/S0104–07072009000100011 Singer, B., & Ryff, C. D. (1999). Hierarchies of life histories and associated health risks. Annals of the New York Academy of Science, 896, 96–115. doi: 10.1111/j.1749– 6632.1999.tb08103.x Singer, B., Ryff, C. D., & Seeman, T. (2004). Operationalizing allostatic load. In J. Schulkin (Ed.) Allostasis, homeostasis, and the costs of psychological adaptation (pp. 113–149). New York, NY: Cambridge University Press. Singer, M. (1996). A dose of drugs, a touch of violence, a case of AIDS: Conceptualizing the SAVA syndemic. Free Inquiry in Creative Sociology, 24(2), 99–110. Slopen, N., Lewis, T. T., Gruenewald, T. L., Mujahid, M. S., Ryff, C. D., Albert, M. A., & Williams, D. R. (2010). Early life adversity and inflammation in African Americans and Whites in the Midlife in the United States Survey. Psychosomatic Medicine, 72, 694–701. doi: 10.1097/PSY.0b013e3181e9c16f Smith, M. A., Makino, S., Kim, S. Y., & Kvetnansky, R. (1995a). Stress increases brainderived neurotropic factor messenger ribonucleic acid in the hypothalamus and pituitary. Endocrinology, 136, 3743–3750. doi: 10.1210/endo.136.9.7649080 Smith, M. A., Makino, S., Kvetnansky, R., & Post, R. M. (1995b). Stress and glucocorticoids affect the expression of brain-derived neurotrophic factor and neurotrophin-3 mRNAs in the hippocampus. Journal of Neuroscience, 15, 1768–1777.

Smith, P. J., Blumenthal, J. A., Hoffman, B. M., Cooper, H., Strauman, T. A., Welsh-Bohmer, K.,…& Sherwood, A. (2010). Aerobic exercise and neurocognitive performance: a metaanalytic review of randomized controlled trials. Psychosomatic Medicine, 72, 239–252. doi: 10.1097/PSY.0b013e3181d14633 Snyder, J. S., Soumier, A., Brewer, M., Pickel, J., & Cameron, H. A. (2011). Adult hippocampal neurogenesis buffers stress responses and depressive behaviour. Nature, 476, 458–461. doi: 10.1038/nature10287 Sole-Padulles, C., Bartres-Faz, D., Junque, C., Vendrell, P., Rami, L., Clemente, I. C.,…, Molinuevo, J. L. (2009). Brain structure and function related to cognitive reserve variables in normal aging, mild cognitive impairment and Alzheimer's disease. Neurobiology of Aging, 30, 1114–1124. doi: 10.1016/j.neurobiolaging.2007.10.008 Stalder, T., Steudte, S., Miller, R., Skoluda, N., Dettenborn, L., & Kirschbaum, C. (2012). Intraindividual stability of hair cortisol concentrations. Psychoneuroendocrinology, 37, 602– 610. doi: 10.1016/j.psyneuen.2011.08.007 Stall, R., Friedman, M., & Catania, J. A. (2008). Interacting epidemics and gay men's health: A theory of syndemics. In R. J. Wolitski, R. Stall, & R. O. Valdiserri (Eds.), Unequal opportunity: Health disparities affecting gay and bisexual men in the United States (pp. 251–274). New York, NY: Oxford University Press. Stall, R., Mills, T. C., Williamson, J., Hart, T., Greenwood, G., Paul, J.,…Catania, J. A. (2003). Association of co-occurring psychosocial health problems and increased vulnerability to HIV/AIDS among urban men who have sex with men. American Journal of Public Health, 93, 939–942. Statistics Canada. (2006a). Aboriginal peoples. 2006 Census of Population. Statistics Canada. https://www12.statcan.gc.ca/census-recensement/2006/rt-td/ap-pa-eng.cfm Statistics Canada. (2006b). Aboriginal peoples—Special Interest Profiles. 2006 Census of Population. Statistics Canada. http://www5.statcan.gc.ca/olc-cel/olc.action?objId=97-564X2006003&objType=46&lang=en&limit=0 Statistics Canada. (2006c). Census Inuit tables. 2006 Census of Population. Statistics Canada. http://www.statcan.gc.ca/pub/89-636-x/89-636-x2008001-eng.htm Statistics Canada. (2008). Life expectancy in Inuit regions of Canada. CANSIM Socioeconomic Database. Statistics Canada. http://www5.statcan.gc.ca/cansim/a26?lang=eng&id=1020706 Staufenbiel, S. M., Penninx, B. W., Spijker, A. T., Elzinga, B. M., & van Rossum, E. F. (2012). Hair cortisol, stress exposure, and mental health in humans: A systematic review. Psychoneuroendocrinology, 38, 1220–1235. doi: 10.1016/j.psyneuen.2012.11.015 Steffener, J., Brickman, A. M., Rakitin, B. C., Gazes, Y., & Stern, Y. (2009). The impact of age-related changes on working memory functional activity. Brain Imaging and Behavior, 3,

142–153. doi: 10.1007/s11682–008–9056-x Steffener, J., Reuben, A., Rakitin, B. C., & Stern, Y. (2011). Supporting performance in the face of age-related neural changes: Testing mechanistic roles of cognitive reserve. Brain Imaging and Behavior, 5, 212–221. doi: 10.1007/s11682–011–9125–4 Steiner, J. L., Murphy, E. A., McClellan, J. L., Carmichael, M. D., & Davis, J. M. (2011). Exercise training increases mitochondrial biogenesis in the brain. Journal of Applied Physiology, 111, 1066–1071. doi: 10.1152/japplphysiol.00343.2011 Sterling, P. (2004). Principles of allostasis: Optimal design, predictive regulation, pathophysiology and rational therapeutics. In J. Schulkin (Ed.) Allostasis, homeostasis, and the costs of adaptation, 629–649. New York, NY: Cambridge University Press. Sterling, P., & Eyer, J. (1988). Allostasis: A new paradigm to explain arousal pathology. In S. Fisher & J. Reason (Eds.), Handbook of life stress, cognition and health (pp. 629–649). New York, NY: Wiley. Stern, Y. (2002). What is cognitive reserve? Theory and research application of the reserve concept. Journal of the International Neuropsychology Society, 8, 448–460. doi: 10.1017.S1355617701020240 Stern, Y. (2006). Cognitive reserve and Alzheimer disease. Alzheimer Disease and Associated Disorders, 20, S69–S74. doi: 10.1016/S1474–4422(12)70191–6 Stern, Y. (2009). Cognitive reserve. Neuropsychologia, 47, 2015–2028. doi: 10.1016/j.neuropsychologia.2009.03.004 Stern, Y. (2012). Cognitive reserve in ageing and Alzheimer's disease. Lancet Neurology, 11, 1006–1012. doi: 10.1016/S1474–4422(12)70191–6 Stern, Y., Albert, S., Tang, M. X., & Tsai, W. Y. (1999). Rate of memory decline in AD is related to education and occupation: Cognitive reserve? Neurology, 53, 1942–1947. doi: 10.1212/WNL.53.9.1942 Stern, Y., Alexander, G. E., Prohovnik, I., & Mayeux, R. (1992). Inverse relationship between education and parietotemporal perfusion deficit in Alzheimers disease. Annals of Neurology, 32, 371–375. Stern, Y., Alexander, G. E., Prohovnik, I., Stricks, L., Link, B., Lennon, M. C., & Mayeux, R. (1995). Relationship between lifetime occupation and, parietal flow—Implications for a reserve against Alzheimers disease pathology. Neurology, 45, 55–60. doi: 10.1097/01.wad.0000213815.20177.19 Stern, Y., Gurland, B., Tatemichi, T. K., Tang, M. X., Wilder, D., & Mayeux, R. (1994). Influence of education and occupation on the incidence of Alzheimer's disease. Journal of the Amercan Medical Association, 271, 1004–1010. doi: 10.1001/jama.1994.03510370056032

Stroud, L. R., Foster, E., Papandonatos, G. D., Handwerger, K., Granger, D. A., Kivlighan, K. T., & Niaura, R. (2009). Stress response and the adolescent transition: performance versus peer rejection stressors. Development and Psychopathology, 21, 47–68. doi: 10.1017/S0954579409000042 Stroud, L. R., Salovey, P., & Epel, E. S. (2002). Sex differences in stress responses: social rejection versus achievement stress. Biological Psychiatry, 52, 318–327. Stumvoll, M., Tataranni, P. A., & Bogardus, C. (2004). The role of glucose allostasis in type 2 diabetes. Reviews in Endocrine and Metabolic Disorders, 5, 99–103. doi: 10.1023/B:REMD.0000021430.56457.2c Sun, J., Wang, S., Zhang, J. Q., & Li, W. (2007). Assessing the cumulative effects of stress: the association between job stress and allostatic load in a large sample of Chinese employees. Work and Stress, 21, 333–347. doi: 10.1080/02678370701742748 Sun, T., Qiao, H., Pan, P. Y., Chen, Y., & Sheng, Z. H. (2013). Motile axonal mitochondria contribute to the variability of presynaptic strength. Cell Reports, 4, 413–419. doi: 10.1016/j.celrep.2013.06.040 Surget, A., Tanti, A., Leonardo, E. D., Laugeray, A., Rainer, Q., Touma, C.,…Belzung, C. (2011). Antidepressants recruit new neurons to improve stress response regulation. Molecular Psychiatry, 16, 1177–1188. doi: 10.1038/mp.2011.48 Suri, D., & Vaidya, V. A. (2012). Glucocorticoid regulation of brain-derived neurotrophic factor: Relevance to hippocampal structural and functional plasticity. Neuroscience, 239, 196– 213. doi: 10.1016/j.neuroscience.2012.08.065 Szanton, S. L., Gill, J. M., & Allen, J. K. (2005). Allostatic load: A mechanism of socioeconomic health disparities? Biological Research in Nursing, 7, 7–15. doi: 10.1016/j.biopsych.2006.03.016 Szyf, M., McGowan, P., & Meaney, M. J. (2008). The social environment and the epigenome. Environmental and Molecular Mutagenesis, 49, 46–60. doi: 10.1002/em.20357 Taivassalo, T., Ayyad, K., & Haller, R. G. (2012). Increased capillaries in mitochondrial myopathy: Implications for the regulation of oxygen delivery. Brain, 135, 53–61. doi: 10.1093/brain/awr293 Taivassalo, T., Jensen, T. D., Kennaway, N., DiMauro, S., Vissing, J., & Haller, R. G. (2003). The spectrum of exercise tolerance in mitochondrial myopathies: a study of 40 patients. Brain 126, 413–423. doi: 10.1093/brain/awg028 Talens, R. P., Boomsma, D. I., Tobi, E. W., Kremer, D., Jukema, J. W., Willemsen, G.,… Heijmans, B. T. (2010). Variation, patterns, and temporal stability of DNA methylation: considerations for epigenetic epidemiology. Federation of American Societies for Experimental Biology Journal, 24, 3135–3144. doi: 10.1096/fj.09–150490

Taliaz, D., Stall, N., Dar, D. E., & Zangen, A. (2010). Knockdown of brain-derived neurotrophic factor in specific brain sites precipitates behaviors associated with depression and reduces neurogenesis. Molecular Psychiatry, 15, 80–92. doi: 10.1038/mp.2009.67 Tamashiro, K. L., Sakai, R. R., Shively, C. A., Karatsoreos, I. N., & Reagan, L. P. (2011). Chronic stress, metabolism, and metabolic syndrome. Stress, 14, 468–474. doi: 10.3109/10253890.2011.606341 Tanti, A., Rainer, Q., Minier, F., Surget, A., & Belzung, C. (2012). Differential environmental regulation of neurogenesis along the septo-temporal axis of the hippocampus. Neuropharmacology, 63, 374–384. doi: 10.1016/j.neuropharm.2012.04.022 Tarullo, A. R., & Gunnar, M. R. (2006). Child maltreatment and the developing HPA axis. Hormones and Behavior 50, 632–639. doi: 10.1016/j.yhbeh.2006.06.010. Tavares, B. C., Barreto, F. A., Lodetti, M. L., Silva, D. M. G. V., & Lessmann, J. C. (2011). Resilience among people with diabetes mellitus. Texto and Contexto Enfermagem, 20, 751– 757. doi: 10.1590/S0104–07072011000400014 Taylor, R. W., & Turnbull, D. M. (2005). Mitochondrial DNA mutations in human disease. Nature Reviews Genetics, 6, 389–402. doi: 10.1038/nrg1606 Taylor, S. E., Klein, L. C., Lewis, B. P., Gruenewald, T. L., Gurung, R. A., & Updegraff, J. A. (2000). Biobehavioral responses to stress in females: tend-and-befriend, not fight-or-flight. Psychological Review, 107, 411–429. doi: 10.1037/0033–295X.107.3.411 Taylor, S. E., Lehman, B. J., Kiefe, C. I., & Seeman, T. E. (2006). Relationship of early life stress and psychological functioning to adult C-reactive protein in the coronary artery risk development in young adults study. Biological Psychiatry, 60, 819–824. doi: 10.1016/j.biopsych.2006.03.016 Tejeda, M. (2006). Nondiscrimination policies and sexual identity disclosure: Do they make a difference in employee outcomes? Employee Responsibilities and Rights Journal, 18, 45–59. Theall, K. P., Drury, S. S., & Shirtcliff, E. A. (2012). Cumulative neighborhood risk of psychosocial stress and allostatic load in adolescents. American Journal of Epidemiology, 176, Suppl. 7, S164–S174. doi: 10.1093/aje/kws185 Thisted, R. A. (2003). Are there social determinants of health and disease? Perspectives in Biological Medicine, 46, S65–S73. doi: 10.1353/pbm.2003.0062 Tomiyama, A. J., O'Donovan, A., Lin, J., Puterman, E., Lazaro, A., Chan, J.,…Epel, E. (2012). Does cellular aging relate to patterns of allostasis? An examination of basal and stress reactive HPA axis activity and telomere length. Physiology and Behavior, 106, 40–45. doi: 10.1016/j.physbeh.2011.11.016 Trejo, J. L., Carro, E., & Torres-Aleman, I. (2001). Circulating insulin-like growth factor I

mediates exercise-induced increases in the number of new neurons in the adult hippocampus. Journal of Neuroscience, 21, 1628–1634. Tremblay, A., & Chaput, J. P. (2012). Obesity: The allostatic load of weight loss dieting. Physiology and Behavior, 106, 16–21. doi: 10.1016/j.physbeh.2011.05.020 Trepanier, L., Juster, R. P., Marin, M. F., Plusquellec, P., Francois, N., Sindi, S.,…Lupien, S. J. (2013). Early menarche predicts increased depressive symptoms and cortisol levels in Quebec girls ages 11 to 13. Development and Psychopathology, 25, 1017–1027. doi: 10.1017/S0954579413000345 Trickett, P. K., Noll, J. G., Susman, E. J., Shenk, C. E., & Putnam, F. W. (2010). Attenuation of cortisol across development for victims of sexual abuse. Development and Psychopathology, 22, 165–175. doi: 10.1017/S0954579409990332 Tsankova, N., Renthal, W., Kumar, A., & Nestler, E. J. (2007). Epigenetic regulation in psychiatric disorders. Nature Reviews Neuroscience, 8, 355–367. doi: 10.1038/nrn2132 Uchino, B. N., Cacioppo, J. T., & Kiecolt-Glaser, J. K. (1996). The relationship between social support and physiological processes: A review with emphasis on underlying mechanisms and implications for health. Psychological Bulletin, 119, 488–531. doi: 10.1037/0033–2909.119.3.488 United Nations. (2003). Millennium Development Goals: A compact among nations to end human poverty. Oxford, UK: Oxford University Press. Vaghri, Z., Guhn, M., Weinberg, J., Grunau, R. E., Yu, W., & Hertzman, C. (2013). Hair cortisol reflects socio-economic factors and hair zinc in preschoolers. Psychoneuroendocrinology, 38, 331–340. doi: 10.1016/j.psyneuen.2012.06.009 Valente, M. M., Bortolotto, V., Cuccurazzu, B., Ubezio, F., Meneghini, V., Francese, M. T.,… Grilli, M. (2012). alpha2delta Ligands act as positive modulators of adult hippocampal neurogenesis and prevent depression-like behavior induced by chronic restraint stress. Molecular Pharmacology, 82, 271–280. doi: 10.1124/mol.112.077636 Valenzuela, M. J., & Sachdev, P. (2006). Brain reserve and dementia: A systematic review. Psychological Medicine, 36, 441–454. doi: 10.1017/S0033291705006264 Valenzuela, M. J., Sachdev, P., Wen, W., Chen, X., & Brodaty, H. (2008). Lifespan mental activity predicts diminished rate of hippocampal atrophy. PLoS One, 3, e2598. doi: 10.1371/journal.pone.0002598 Valera, B., Dewailly, E., & Poirier, P. (2008). Cardiac autonomic activity and blood pressure among Nunavik Inuit adults exposed to environmental mercury: A cross-sectional study. Environmental Health, 7, 29. doi: 10.1186/1476–069X-7–29 Valera, B., Dewailly, E., & Poirier, P. (2009). Environmental mercury exposure and blood

pressure among Nunavik Inuit adults. Hypertension, 54, 981–986. doi: 10.1161/HYPERTENSIONAHA.109.135046 Valera, B., Dewailly, E., & Poirier, P. (2013). Association between methylmercury and cardiovascular risk factors in a native population of Quebec (Canada): A retrospective evaluation. Environmental Research, 120, 102–108. doi: 10.1016/j.envres.2012.08.002 Valera, B., Muckle, G., Poirier, P., Jacobson, S. W., Jacobson, J. L., & Dewailly, E. (2012). Cardiac autonomic activity and blood pressure among Inuit children exposed to mercury. Neurotoxicology, 33, 1067–1074. doi: 10.1016/j.neuro.2012.05.005 Van Cauter, E., & Spiegel, K. (1999). Sleep as a mediator of the relationship between socioeconomic status and health: A hypothesis. Annals of the New York Academy of Science, 896, 254–261. doi: 10.1111/j.1749–6632.1999.tb08120.x Veena, J., Srikumar, B. N., Mahati, K., Bhagya, V., Raju, T. R., & Shankaranarayana Rao, B. S. (2009). Enriched environment restores hippocampal cell proliferation and ameliorates cognitive deficits in chronically stressed rats. Journal of Neuroscience Research, 87, 831– 843. doi: 10.1002/jnr.21907 Verner, M. A., Plusquellec, P., Muckle, G., Ayotte, P., Dewailly, E., Jacobson, S. W.,… Haddad, S. (2010). Alteration of infant attention and activity by polychlorinated biphenyls: Unravelling critical windows of susceptibility using physiologically based pharmacokinetic modeling. Neurotoxicology, 31, 424–431. doi: 10.1016/j.neuro.2010.05.011 Vivian, J. A., & Miczek, K. A. (1999). Interactions between social stress and morphine in the periaqueductal gray: Effects on affective vocal and reflexive pain responses in rats. Psychopharmacology (Berlin), 146, 153–161. doi: 10.1007/s002130051101 Vos, M., Lauwers, E., & Verstreken, P. (2010). Synaptic mitochondria in synaptic transmission and organization of vesicle pools in health and disease. Frontiers in Synaptic Neuroscience, 2, 139. doi: 10.3389/fnsyn.2010.00139 Vos, T., Barker, B., Begg, S., Stanley, L., & Lopez, A. D. (2009). Burden of disease and injury in Aboriginal and Torres Strait Islander peoples: The Indigenous health gap. International Journal of Epidemiology, 38, 470–477. doi: 10.1093/ije/dyn240 Vyskocil, A., Fiala, Z., Ettlerova, E., & Tenjnorova, I. (1990). Influence of chronic lead exposure on hormone levels in developing rats. Journal of Applied Toxicology, 10, 301–302. Waldo, C. R. (1999). Working in a majority context: A structural model of heterosexism as minority stress in the workplace. Journal of Couseling Psychology, 46, 218–232. doi: 10.1037/0022–0167.46.2.218 Wallace, D. C. (2005). A mitochondrial paradigm of metabolic and degenerative diseases, aging, and cancer: A dawn for evolutionary medicine. Annual Review of Genetics, 39, 359– 407. doi: 10.1146/annurev.genet.39.110304.095751

Wallace, D. C. (2010). Colloquium paper: Bioenergetics, the origins of complexity, and the ascent of man. Proceedings of the National Academy of Sciences of the USA, 107, Suppl. 2, 8947–8953. doi: 10.1073/pnas.0914635107 Wallace, D. C. (2013). A mitochondrial bioenergetic etiology of disease. Journal of Clinical Investigation, 123, 1405–1412. doi: 10.1172/JCI61398 Wang, H. X., Karp, A., Winblad, B., & Fratiglioni, L. (2002). Late-life engagement in social and leisure activities is associated with a decreased risk of dementia: A longitudinal study from the Kungsholmen project. American Journal of Epidemiology, 155, 1081–1087. doi: 10.1093/aje/155.12.1081 Wang, J. W., David, D. J., Monckton, J. E., Battaglia, F., & Hen, R. (2008). Chronic fluoxetine stimulates maturation and synaptic plasticity of adult-born hippocampal granule cells. Journal of Neuroscience, 28, 1374–1384. doi: 10.1523/JNEUROSCI.3632–07.2008 Webb, S. (2009). Palaeopathology of Aboriginal Australians: Health and disease across a hunter-gatherer continent. Cambridge, UK: Cambridge University Press. Weber, K., Rockstroh, B., Borgelt, J., Awiszus, B., Popov, T., Hoffmann, K.,…Propster, K. (2008). Stress load during childhood affects psychopathology in psychiatric patients. BMC Psychiatry, 8, 63. doi: 10.1186/1471–244X-8–63 Weinstein, M., Goldman, N., Hedley, A., Yu-Hsuan, L., & Seeman, T. (2003). Social linkages to biological markers of health among the elderly. Journal of Biosocial Science, 35, 433–453. doi: 10.1017/S0021932003004334 Westerlund, H., Gustafsson, P. E., Theorell, T., Janlert, U., & Hammarstrom, A. (2013). Parental academic involvement in adolescence, academic achievement over the life course and allostatic load in middle age: A prospective population-based cohort study. Journal of Epidemiology and Community Health, 67, 508–513. doi: 10.1136/jech-2012–202052 Whalley, L. J., Deary, I. J., Appleton, C. L., & Starr, J. M. (2004). Cognitive reserve and the neurobiology of cognitive aging. Ageing Research Reviews, 3, 369–382. doi: 10.1016/J.Arr.2004.05.001 Widom, C. S., DuMont, K., & Czaja, S. J. (2007). A prospective investigation of major depressive disorder and comorbidity in abused and neglected children grown up. Archives of General Psychiatry, 64, 49–56. doi: 10.1001/archpsyc.64.1.49 Wilkinson, R. G. (1996). Unhealthy societies: The affliction of inequality. London, UK: Routledge. Williams, D. R., Gonzalez, H. M., Williams, S., Mohammed, S. A., Moomal, H., & Stein, D. J. (2008). Perceived discrimination, race and health in South Africa. Social Science and Medicine, 67, 441–452. doi: 10.1016/j.socscimed.2008.03.021

Wilson, R. S., Beckett, L. A., Barnes, L. L., Schneider, J. A., Bach, J., Evans, D. A., & Bennett, D. A. (2002a). Individual differences in rates of change in cognitive abilities of older persons. Psychology of Aging, 17, 179–193. doi: 10.1037/0882–7974.17.2.179 Wilson, R. S., Mendes De Leon, C. F., Barnes, L. L., Schneider, J. A., Bienias, J. L.,…Bennett, D. A. (2002b). Participation in cognitively stimulating activities and risk of incident Alzheimer disease. Journal of the Amercan Medical Association, 287, 742–748. doi: 10.1001/jama.287.6.742 Wong, E. Y., & Herbert, J. (2004). The corticoid environment: a determining factor for neural progenitors' survival in the adult hippocampus. European Journal of Neuroscience, 20, 2491– 2498. doi: 10.1111/j.1460–9568.2004.03717.x World Health Organization. (2006). Gender equality, work and health: A review of the evidence. Geneva, Switzerland: Author. http://www.who.int/gender/documents/Genderworkhealth.pdf World Health Organization. (2011). American Samoa. WHO Western Pacific Country Health Information Profiles. Geneva, Switzerland. http://www.wpro.who.int/countries/asm/en/ World Health Organization. (2012). Global health indicators. World Health Statistics 2012. Geneva, Switzerland. http://www.who.int/gho/publications/world_health_statistics/2012/en/ Yamada, J., Stevens, B., de Silva, N., Gibbins, S., Beyene, J., Taddio, A.,…Koren, G. (2007). Hair cortisol as a potential biologic marker of chronic stress in hospitalized neonates. Neonatology, 92, 42–49. doi: 10.1159/000100085 Yang, S. H., Liu, R., Perez, E. J., Wen, Y., Stevens, S. M., Jr., Valencia, T.,…Simpkins, J. W. (2004). Mitochondrial localization of estrogen receptor beta. Proceedings from the National Academy of Sciences of the USA, 101, 4130–4135. doi: 10.1073/pnas.0306948101 Yu, T., Robotham, J. L., & Yoon, Y. (2006). Increased production of reactive oxygen species in hyperglycemic conditions requires dynamic change of mitochondrial morphology. Proceedings of the National Academy of Sciences of the USA, 103, 2653–2658. doi: 10.1073/pnas.0511154103 Zeiders, K. H., Doane, L. D., & Roosa, M. W. (2012). Perceived discrimination and diurnal cortisol: examining relations among Mexican American adolescents. Hormones and Behavior, 61, 541–548. doi: 10.1016/j.yhbeh.2012.01.018 Zota, A. R., Shenassa, E. D., & Morello-Frosch, R. (2013). Allostatic load amplifies the effect of blood lead levels on elevated blood pressure among middle-aged U.S. adults: A crosssectional study. Environmental Health, 12, 64. doi: 10.1186/1476–069X-12–64 1 Color versions of Figure 8.1 are available at

http://onlinelibrary.wiley.com/book/10.1002/9781119125556

Chapter 9 Competence and Psychopathology in Development Keith B. Burt, J. Douglas Coatsworth, and Ann S. Masten Acknowledgments: We thank Glenn I. Roisman and Michael F. Troy for comments on a prior version of this chapter. DEFINING COMPETENCE AND PSYCHOPATHOLOGY: HISTORICAL LEGACIES AND THEORETICAL WORK RELATIONS BETWEEN COMPETENCE AND PSYCHOPATHOLOGY IMPAIRMENT AND MENTAL DISORDER: SYSTEMS, MEASURES, AND RESEARCH SUPPORT Formal Systems and Instruments for Assessing and Classifying Impairment Separating Impairment from Definitions of Mental Disorder Associations Between Psychopathology and Functional Impairment: Empirical Evidence NIMH Research Domain Criteria Project Impairment: Conclusions CASCADE MODELS Conceptual and Statistical Background Empirical Findings Cascade Models: Summary INTERVENTIONS TO PROMOTE COMPETENCE AND REDUCE PSYCHOPATHOLOGY Conceptual Models Overview of Interventions Targeting Improvements in Core Competencies Promotion and Prevention by Building Nurturing Environments Testing Mediating Processes Moderation of Prevention Effects via Gene–Environment Interaction Prevention Programs: Summary CONCLUSIONS AND FUTURE DIRECTIONS REFERENCES

A core theoretical feature of developmental psychopathology is the simultaneous consideration of normal and abnormal development; this is historically prominent and ongoing to the present day, shaping research agendas for the future (Cicchetti, 1984, 2013; Rutter & Sroufe, 2000; Sroufe & Rutter, 1984). The developmental psychopathology perspective comprises several tenets that guide investigation, including a focus on processes that unfold over time, a consideration of interactions across multiple levels of analysis, and an appreciation for diverse developmental pathways through concepts of equifinality and multifinality (Cicchetti, 2006, 2013; Cicchetti & Rogosch, 1996; Masten, 2006; Pickles & Hill, 2006). Based on these foundational tenets, in this chapter we review the interrelations of adaptive and maladaptive functioning, broadly considered. Rather than focus on one specific mental disorder or diagnostic category, we detail theory and evidence on associations among prominent domains of competence (e.g., academics, social relationships) and broad dimensions of psychopathology (e.g., internalizing and externalizing symptoms) in youth. Several specific diagnostic categories of psychopathology are considered in the process. Overall, we aim to provide a review of empirical work detailing the nature of these associations as well as theoretical and methodological discussion of how one might interpret the empirical findings. The chapter is divided into three main sections. First, we briefly review historical and theoretical discussions of the concept of competence, particularly from a developmental tasks framework. Second, in the central part of the chapter, we discuss empirical reports on the links between competence and psychopathology, working from two main traditions: (1) classification systems for mental illness, in which competence can be considered the opposite of functional impairment, and (2) statistical models relating competence and psychopathology, with attention to details of analytic methodology as well as common substantive findings. The latter section focuses particularly on a class of longitudinal mediation models termed cascade models (Masten & Cicchetti, 2010). Finally, we consider the literature on competencepromoting prevention and intervention programs, with a prominent focus on results from randomized controlled trials, which has the potential to provide both crucial practical benefits to youth as well as rigorous testing of causal claims.

Defining Competence and Psychopathology: Historical Legacies and Theoretical Work Here we briefly recapitulate selected key developments in the evolution of thinking on competence and psychopathology; for more detailed discussions, please see the two prior editions of this chapter (Masten, Burt, & Coatsworth, 2006; Masten & Coatsworth, 1995). Definitions of psychopathology have evolved considerably since the mid-twentieth-century advent of the Diagnostic and Statistical Manual of the American Psychiatric Association (the DSM system), but debates continue on the nature of mental disorder, in particular to what extent mental illness can be conceptualized as categorically versus continuously distributed in the population (Coghill & Sonuga-Barke, 2012; Pickles & Angold, 2003) and how best to assess the core features of disorder, especially for some youth whose problems may be more distressing to family members or school personnel than to themselves. We consider definitional

issues of mental disorder further later. We have previously presented a working definition of competence as “a family of constructs related to the capacity or motivation for, process of, or outcomes of effective adaptation in the environment, often inferred from a track record of effectiveness in age-salient developmental tasks, and always embedded in developmental, cultural, and historical context” (Masten et al., 2006, p. 704). This definition is indebted to historical work on developmental tasks: the adaptive, age-related criteria by which cultures, families, and individuals judge how well development is proceeding (see McCormick et al., 2011). The concept of developmental tasks was advanced significantly by educational theorist Robert Havighurst's (1972) work on the idea of a progression of achievements that children, adolescents, and adults are expected to accomplish by members of the broader society. Developmental task theories are in turn based partly on (a) earlier psychoanalytic work on the resolution or persistence of psychosocial crises in an individual progression through specific stages (Erikson, 1968); and (b) work on ego psychology and personality development that focused on the integration of cognition, affect, and behavior (Loevinger, 1966) and constructs of ego control and ego resiliency (Block & Block, 1980). The latter two constructs describe not only the effective modulation of impulses but also the flexible control of that modulation. In addition, the history of developmental tasks research owes a substantial debt to work on intellectual functioning in youth, which has had a strong focus on assessing the ability of children to act effectively in their particular environmental context (Binet & Simon, 1905; Charlesworth, 1979; Sternberg, 1985; Wechsler, 1958). Although intelligence (the inferred ability) and “intelligent behavior” have been closely linked to other domains of competence in theory and research, they often are distinguished from other competence domains in assessment and diagnosis. For example, in diagnosis of intellectual disability, the psychometric assessment of intellect is joined with a judgment of adaptive behavior, which often includes practical or social skills from a formal assessment, such as the Vineland Adaptive Behavior Scales (VABS; Sparrow, Cicchetti, & Balla, 2005). Finally, researchers interested in the development of self-concept and mastery motivation have developed theories on the intrinsic motivation of humans to master their environment (White, 1959), building on evolutionary and biological foundations to integrate cognitive and behavioral constructs such as self-perception, locus of control, and self-efficacy (Bandura, 1997; Harter, 1978). This work has advanced definitions of competence by focusing attention on the interrelations between internal judgments of effectiveness across domains as well as judgments made by others. Each of these areas has informed the study of developmental tasks. Havighurst (1972) proposed that specific age-salient tasks form the basis by which members of society judge whether or not an individual's developmental trajectory has been successful. His initial series of specific tasks was grounded in middle-class mid-twentieth-century American values. For example, at earlier ages in development, these tasks included learning to communicate and distinguishing right from wrong; at older ages, developmental tasks included academic success, forming close relationships, and civic responsibility. Other writers have proposed

organized sequences of developmental challenges and tasks, in which successful achievement of prior tasks set the stage for those coming later (Greenspan, 1981; Sroufe, 1979; Waters & Sroufe, 1983). More recent examples of developmental tasks in contemporary industrialized societies organized by developmental period can be found in several of our collaborative publications (Masten & Coatsworth, 1998; Masten et al., 2006; McCormick et al., 2011). In similar historical and cultural contexts, independent investigators often identify similar developmental tasks. Nonetheless, developmental task theorists have posited that there also would be some tasks unique to a particular context of development, reflecting the values and requirements for success in different cultures, ecologies, or periods of history (Masten et al., 2006; McCormick et al., 2011). Over the decades, there also have been significant debates about the definition and assessment of competence. Salient points of debate include whether to include internal adaptation (e.g., subjective well-being) as well as external adaptation (relationship-based or otherwise) in criteria for competent functioning; to what extent processes versus individual attributes are highlighted; and, crucially, how broad or narrow each domain of competence should be (see Masten et al., 2006; McCormick et al., 2011). For example, some theories about developmental tasks view identity as a key task (e.g., Erikson, 1968) or sequence of tasks (see Motti-Stefanidi, 2015) whereas others focus on meeting criteria for expected behavior in the environment (such as schoolwork or work achievement, friendship, or law-abiding conduct). Finally, we emphasize that the study of positive development and the nature and measurement of developmental tasks is an active area of research. For example, international work to develop a core set of child well-being indicators is ongoing. One such effort is UNICEF's Innocenti Report Card 7, derived from the United Nations Convention on the Rights of the Child and including domains of material well-being, health and safety, education, peer and family relationships, subjective well-being, and behavior and risk (United Nations Children's Fund, 2007). More generally, Mahoney and Bergman (2002) have discussed conceptual and statistical aspects of positive development work from a holistic and person-centered perspective. Much of their discussion dovetails with the work described earlier, although they provide a helpful reminder that both developmental tasks themselves and the processes by which youth may attain (or fail to attain) them often vary by context (e.g., high- versus low-risk youth). Mahoney and Bergman write from the vantage point of positive psychology and thus provide connections from this work to other (largely adult-based) research on challenging but crucial constructs such as optimism, wisdom, and agency.

Relations between Competence and Psychopathology There are several reasons to expect consistent associations between measures of competence and measures of psychopathology. Unfortunately, there is a continuing and persistent risk of misinterpreting associations, in part because of the conceptual and methodological overlaps in the definitions and operationalization of these concepts. First, the concepts themselves may overlap, as when certain diagnostic categories (e.g., conduct disorder in the DSM system) share defining criteria with certain developmental tasks (rule-abiding conduct), a problem

relevant to debates on impairment and mental disorder discussed later. Although not of great interest theoretically, this permeable boundary between constructs has the potential to mislead research if one is not extremely careful about operational definitions. In addition, even constructs that are nominally discrete may share overlapping assessment criteria, a problem that is well known in research on associations between temperament and psychopathology and has led to calls for revised assessed instruments in that area (Lahey, 2004). As noted in our prior work (Masten et al., 2006), many common measures of functional impairment contain content closely related to content of competence measurements, albeit with a focus on a more negative range of functioning. Thus, to the extent that impairment might be considered the purview of mental disorder, these concepts will be confounded. If overlapping definitions and shared assessment content can be convincingly ruled out or minimized, research focus shifts to three non-mutually exclusive categories well-known for explaining an association between two constructs A and B: Either A causes B, B causes A, or one or more additional constructs cause both A and B. This presentation of competencepsychopathology relations is somewhat reductionistic, in that processes underlying these directional, causal effects likely operate simultaneously and continuously, consistent with the nature of dynamic systems (Gottlieb, 2007; R. M. Lerner, 2002; Overton, 2013). Nonetheless, we feel that the delineation of these competing explanations is of value, as it can help identify areas for further conceptual and methodological/statistical work. For the constructs considered here, this logic means that successes or failures in competence cause changes in psychopathological symptoms (improvements or exacerbations), that symptoms (or their immediate effects) alter subsequent competence, or that one or more of a series of third variables operate to cause changes both in competence and behavioral or emotional problems. Importantly, in this chapter, we usually discuss the directional causal hypotheses in terms of failures in competence contributing to psychopathology or psychopathology undermining adaptive competence. This is because the majority of theory and research in this area is conceptualized in terms of failure. However, it should be kept in mind throughout that associations of the continuous, dimensional variables that index levels of competence or psychopathology are also consistent with the possibility that competence successes prevent the development or exacerbation of future symptoms or that improvement in symptoms leads to greater subsequent competence (Masten, 2001; Masten et al., 2006). Potential third-variable explanations between competence and symptoms can take several forms. Masten and colleagues (2006) detail a number of examples that show how individual difference variables (e.g., intelligence and personality traits such as impulsivity, introversion, and conscientiousness) and other contextual variables (e.g., received parenting and child maltreatment) might serve as common causes of both psychopathology and competence. Central to this discussion is the idea of cumulative risk and risk gradients: the well-documented finding of an increased likelihood of problems across multiple domains with increased aggregate number of risk factors (Masten, Morison, Pelligrini, & Tellegen, 1990; Rutter, 1979; Sameroff, 2000; see also Obradović, Shaffer, & Masten, 2012). Although delineation of common-cause variables is clearly of importance, there are also

ambiguities in interpretation of the statistical finding when a third variable accounts for the association between two others. As a simple but likely fairly common example, there may be situations in which the third variable serves as a proxy for a prior assessment of one of the two focal variables, complicating inferences about direction of effects (see also Card & Barnett, 2015). Even in situations where such ambiguity is clearly implausible, a larger (and nearubiquitous) concern is that it is in practice impossible to include all relevant common-cause variables in a given analysis. This fact makes the theory-driven selection of key covariates of primary importance, as well as the integration of results across studies and the replication of key results with experimental evidence (from either randomized prevention/intervention studies or natural experiments) wherever possible. Helena Kraemer and her colleagues (1997) have attempted to clarify this picture with reference to risk factors by generating a taxonomy that distinguishes causal risk factors from fixed or variable markers (noncausal factors) as well as a more differentiated system describing how risk factors might work together as mediators, moderators, or proxies to affect an outcome of interest (Kraemer, Stice, Kazdin, Offord, & Kupfer, 2001). Of course, these three primary explanations for competence-symptom associations can operate in tandem, which makes testing of unique predictions that convincingly elevate one or more of the potential reasons for competence-psychopathology co-occurrence extremely difficult. Developmental theorists have strengths in recognizing the complexity of forces at work in this arena: In particular, as noted, writers in this area have specified broader theoretical orientations that explicitly acknowledge the bidirectional nature of the relation between competence and psychopathology and the importance of context, at multiple levels, in moderating that association (e.g., Mahoney & Bergman, 2002; Overton, 2013; Sameroff, 2010). But a continuing challenge for the field is the translation of such theories into testable and falsifiable predictions that can begin to rule out one or more possible explanations for observed associations, despite explicit acknowledgment of the interplay between complex theories and relatively new statistical methods (Jelicic, Theokas, Phelps, & Lerner, 2007). In general, some of the theories that have been most influential in guiding competencepsychopathology work have been more specific, aiming to lay out in detail the processes by which a specific failure of competence is connected with a specific aspect of psychopathology. One example of a well-developed framework for competence failures contributing to subsequent psychopathology is the dual failure model proposed by Patterson, Capaldi, and colleagues, which posits that failures in the key developmental tasks of academics and social relationships/peer acceptance lead to increased depressive symptoms (Capaldi, 1992; Patterson & Stoolmiller, 1991). This model, supported by well-controlled longitudinal evidence, suggests that early behavior problems and family conflict can lead to disrupted school achievement and spreading of negative interaction patterns to peers, which in turn appear to result in increased dysphoric mood. Cole and colleagues have proposed a similar direction of effect between social failure and increased depressive symptoms, one that is mediated by perceived competence (Cole, Martin, Powers, & Truglio, 1997; Tram & Cole, 2000). Cascade modeling work from each of these research teams is reviewed in more detail in a subsequent section. Importantly, these theories have proved generative for diverse other

work: for example, data from the male Developmental Trends Study informed by the failure model has suggested that oppositional defiant disorder (ODD) predicted both later conduct disorder (CD) and also later affective disorder, while CD was related to later affective disorder indirectly via functional impairment (Burke, Loeber, Lahey, & Rathouz, 2005). The second edition of this chapter presented several additional examples of possible competence failures contributing to increases in symptoms of psychopathology as well as the reverse—scenarios in which clearly defined mental illness leads directly or indirectly to problems in age-salient developmental tasks (Masten et al., 2006). As a group, these examples represent a wide array of disorders, tasks, and problems, with ample initial evidence for effects in both directions. It is important to recognize that a search for a single, universal mechanism of action for these effects might be misguided. Indeed, the cascade work reviewed later in the chapter has dealt with this complexity of constructs by limiting primary focus to the key competence domains of academics and social competence—the latter itself a complex construct containing facets of self-concept, peer acceptance and reputation, victimization and bullying, and close friendship—and the broad problem domains of internalizing and externalizing. However, selected papers have expanded the methodology to other constructs, as detailed further later. We now turn to a more detailed consideration of work linking competence and psychopathology through the broad areas of impairment in mental disorder classification systems and longitudinal cascade models. Although mutually informing, these two bodies of work have been conducted by separate research teams working from somewhat different theoretical perspectives and using different samples and methods. We focus primarily on work published since the prior edition of this chapter in 2006.

Impairment and Mental Disorder: Systems, Measures, and Research Support In the second edition of this chapter, we highlighted how competence has been incorporated into the major diagnostic systems for psychiatry and psychology, which is primarily through the opposite of competence, namely functional impairment (Masten et al., 2006). We also discussed Jerome Wakefield's (1997) harmful dysfunction analysis of the concept of mental disorder, which has been influential in some of the philosophical work underpinning prior revision cycles of the DSM. Although the revision process leading to the current version, the fifth edition of the Diagnostic and Statistical Manual of Mental Disorders (DSM-5) (American Psychiatric Association, 2013), included discussion of the nature and definition of mental disorder itself, the issue of impairment in defining mental disorder has been relatively uncontroversial. For example, minor wording changes to the global definition of mental disorder were added to explicitly describe distress and/or impairment as a consequence of disorder, which was implicitly assumed to be the case in prior editions (First & Wakefield, 2010). At the same time, the topic of impairment was raised as an important issue for working groups

to consider when revising disorder criteria (C. Lord, personal communication, February 2013). Partly to align the system more closely with the International Classification of Diseases (ICD) system in use throughout much of the rest of world, DSM working groups were charged with considering the implications of separating out functional impairment from the criteria sets themselves. However, it appears that this task was extremely difficult, as many criteria sets are difficult to measure without some reference to impairment, especially those related to social functioning (e.g., multiple criteria for autism spectrum disorder), and DSM-5 has retained the clinical significance criterion while acknowledging international efforts to separate symptoms from impairment (American Psychiatric Association, 2013). Despite this situation, concerns have continued to be raised regarding the confounding of symptoms and impairment across multiple aspects of DSM, including some criteria sets (especially in the externalizing domain). The collapsing of symptom severity and functional impairment into a single numeric score for Axis V, the Global Assessment of Functioning (GAF), received particular criticism—so much so that DSM-5 revamped the entire multiaxial system, in the process replacing the GAF with the World Health Organization (WHO) Disability Assessment Schedule—Second Edition (WHODAS 2.0, detailed later). However, some researchers continue to explore the SOFAS (Social and Occupational Functioning Assessment Scale) and GARF (Global Assessment of Relational Functioning) scales from prior editions, which are the closest cousins to the GAF. For example, Smith and colleagues (2011) showed in a sample of psychotic patients that symptom severity ratings contributed to substantially more of the variance in GAF scores than did impairment, largely due to rating conventions that set a ceiling for functioning scores in the presence of psychosis. However, database searches indicate that SOFAS research continues to consist primarily of relatively modest-size intervention studies that employ the scale to track patient progress. Although valuable, this work does not provide data for critical evaluation of competing theoretical frameworks for impairment. Turning specifically to youth disorders, a review by Rapee, Bögels, van der Sluis, Craske, and Ollendick (2012) gave careful consideration to a number of important conceptual issues related to impairment. To begin with, the authors noted that aspects of context, culture, gender, and developmental stage will influence the nature and extent of functional impairment. For example, parents or other family members may consciously or unconsciously “accommodate” child behavioral or emotional problems in such a manner that the current functional impact is limited, with either adaptive or maladaptive consequences for later development. In addition, Rapee and colleagues detailed the myriad ways in which impairment has been assessed, from single-score global measures to highly differentiated tools. The latter are used mainly in research settings or in specialty clinics, and a subset of these measures was reviewed in the second edition of this chapter (for additional reviews, see Kutash, Lynn, & Burns, 2008; Winters, Collett, & Myers, 2005). This topic is highly relevant for our broader discussion of competence and psychopathology, because it highlights how researchers and theorists working from a mental disorder classification standpoint have measured impairment. There is reasonable agreement that, for youth, areas of social functioning (including family relationships), academics (and work for adolescents in some contexts), and self-care/leisure

activities are primary measurement domains for impairment. For example, Bird and colleagues (2005) developed the 23-item Brief Impairment Scale (BIS) as an impairment measure that maintains not only the reliability and fast administration of scales such as the GAF or Children's Global Assessment Scale (CGAS; Shaffer et al., 1983) but also the breakdown of impairment across core domains that is characteristic of more differentiated scales; the items are grouped into interpersonal relations, school/work functioning, and self-fulfillment, which includes hobbies and other activities. Rapee and colleagues (2012) also detail the challenges inherent in impairment research in youth as compared with adult populations. The issue of multiple informants arises in a related fashion as it does for symptom assessment: It is sometimes unclear who is the most reliable or valid informant on impairment and, when information gathered from multiple informants is discrepant, how to best aggregate it (Sanford, Offord, Boyle, Peace, & Racine, 1992). This issue is of particular concern given evidence that, in adults, the diagnostically reported “clinically significant impairment” criterion does not always map onto objective measures of impairment, such as work days lost (Breslau & Alvarado, 2007). Additionally, for youth who are assigned more than one mental disorder diagnosis, it is often extremely difficult to disentangle which aspects of impairment are consequences of which disorder, independent of broader concerns regarding the validity of specific pairs of comorbid diagnoses (Angold, Costello, & Erkanli, 1999a; Sroufe, 1997). Thus, issues of impairment connect with broader critiques of DSM and proposals to move the field toward mixed categorical-dimensional systems of taxonomy (Drabick, 2009; Yates, Burt, & Troy, 2011), which may require different systems across different classes of disorders (see Coghill & Sonuga-Barke, 2012, for relevant evidence and discussion, as well as the Research Domain Criteria initiative discussed later). Despite thoughtful reviews such as that by Rapee and colleagues (2012), research on impairment continues to be fragmented, due in part to the diverse stakeholders in classification systems and the significant implications for “caseness” represented by different uses of impairment. For example, Leclair, Leclair, and Brigham (2009) detailed the complications of recent changes to the American Medical Association's process for rating medical impairment for determination of Social Security disability status in the United States and the difficulties of how to reliably and validly apply these standards to mental illness. In addition, distinctions between symptoms and disability have implications for community care, since disabilityrelated care is often provided by organizations other than traditional mental health services (Whiteford, 2009). To help advance this discussion, we consider some specific taxonomic systems of impairment and new dimensional approaches for mental disorder research in more detail as well as recent work on the association of disorder symptoms with impairment.

Formal Systems and Instruments for Assessing and Classifying Impairment The study of impairment has been complicated by the existence of several closely related classification systems that have arisen for specific purposes. One of the most influential is the International Classification of Functioning, Disability, and Health (ICF; WHO, 2001). The

ICF is a wide-ranging classification system designed to categorize the various ways in which both physical and mental illness can impact an individual's life: major groupings of this system are (1) bodily functions and structures, (2) activities, and (3) participation, with allowance for coding of external environmental factors and an overarching biopsychosocial conceptual model of disability (Kostanjsek, 2011). The domains of impairment most relevant for this chapter fall into the ICF's activities and participation categories, which contain some entries similar to widely studied competence domains as well as others that are less often discussed. For example, detailed aspects of interpersonal relationships can be coded, as well as learning and educational skills, various modes of communication, civic engagement, formal education and employment. Consistent with a dual focus on physical and mental health, the ICF also contains detailed codes related to mobility and self-care. A measurement tool referenced earlier, the WHO's Disability Assessment Schedule—Second Edition (WHODAS 2.0) has been developed to assess key domains from the ICF (Üstün et al., 2010). The WHODAS 2.0 was refined via factor analysis and extensively field-tested globally and has been translated into 27 languages. Although items load on a common disability factor, content is also divided into six categories: cognition, mobility, self-care, getting along with others, life activities (e.g., school, work, and household tasks), and community participation. Overall, the WHODAS 2.0 is an impressive undertaking, albeit one developed and validated solely on adult samples; study of its potential application to diverse child samples is more recent (e.g., Scorza et al., 2013). As described earlier, it has replaced the GAF for assessment of impairment in DSM-5. Although the ICF has been an impressive attempt to provide a universal categorization of impairment applicable across all types of health/illness, in many respects it is not developmental in nature, with implicit adult norms and frames of reference. Therefore, it is promising to see the development of the International Classification of Functioning, Disability, and Health–Children and Youth Version (ICF-CY), a closely related classification system of impairment and disability focused on children and adolescents (WHO, 2007). The ICF-CY adds codes for items such as babbling, accepting novelty and adapting one's activity level that may be more applicable to youth; however, most items remain identical to the adult ICF, and developmental norms are not discussed. Although discrete assessment measures directly based on the ICF-CY have to our knowledge not yet been developed, recent research has linked existing measures of adaptive behavior to ICF-CY codes. For example, a content analysis of one widely used scale of adaptive behavior in youth suggests that most scale items were reasonably well captured by the codes (Gleason & Coster, 2012). In addition, work attempting to link assessment measures to the ICF-CY is ongoing (e.g., Castro, Ferreira, Dababnah, & Pinto, 2013; Gan, Tung, Yeh, & Wang, 2013), as is work building consensus on the subsets of ICF-CY elements most relevant to ADHD (Bölte et al., 2014a; de Schipper et al., 2015) and autism spectrum disorder (Bölte et al., 2014b). Others have also attempted to link adaptive behavior to impairment and disability within narrower domains of disorder. Writing mainly from a background in intellectual disability and autism spectrum disorder research, Ditterline and Oakland (2009) provided a comprehensive review of relations between the constructs of adaptive behavior and impairment, as experts in

these domains have been grappling with effective ways of classifying impairment for some time. As detailed by Ditterline and Oakland, the American Association on Intellectual and Developmental Disabilities (AAIDD) groups adaptive skills into ten narrowband domains: communication, community use, functional academics, home/school living, health and safety, leisure, self-care, self-direction, social (skills), work (skills), and motor skills (AAIDD/AAMR, 2002). These 10 areas are grouped in turn into three broad categories: conceptual skills, social skills, and practical skills. Ditterline and Oakland also briefly reviewed well-validated measures of adaptive functioning, such as the Scales of Independent Behavior-Revised (SIB-R; Bruininks, Woodcock, Weatherman, & Hill, 1996) and the VABS (Sparrow et al., 2005), designed to tap into some or all of the above-mentioned areas. The clear overlap among the constructs and domains highlighted by the AAIDD and by the WHO in their respective systems is evident from comparing these lists. Some investigators have tried to capture a middle ground between extremely brief impairment measures, such as the GAF, and more extensive measures, such as the SIB-R or VABS. Whiteside (2009) described a youth adaptation of the widely used Sheehan Disability Scale designed to tap child and parent social, educational/occupational, and family functioning. Initial validation results showed that this measure contributed unique predictive power beyond symptom severity for the outcome of presentation for treatment. In addition, a promising report by van der Knaap and colleagues (2009) described the development of the Adolescents' Tasks and Skills Questionnaire (TASQ), designed to assess impairment in youth living in residential facilities. This measure is noteworthy for its rare focus on positive outcomes in a clinical setting with seriously disturbed youth. The broad dimensions of the TASQ include peer relations, autonomy and self-management, school competence, job competence, romantic competence, and self-care. Initial validation shows expected negative associations between competence domains and behavioral and emotional problems; further validity evidence, including incremental validity for future outcomes, is needed. Finally, recent work by Russell Barkley, focused primarily on impairment resulting from attention disorders, has argued strongly for a statistical-normative conceptualization of impairment. Thus, for both youth and adults, individuals are compared to age- and genderbased norms on a series of impairment areas. Barkley has developed impairment scales for adults (the Barkley Functional Impairment Scale [BFIS]; Barkley, 2012a) and youth (Barkley Functional Impairment Scale for Children and Adolescents [BFIS-CA]; Barkley, 2012b) that show promise as a middle ground between GAF ratings and more time-consuming instruments. For adults, the BFIS includes domains such as home and work responsibilities, social relationships, finances, sexual relationships, and self-care. For youth, the BFIS-CA includes many domains already discussed, including friendships, schoolwork, chores, and family relationships. These measures have already proven useful in differentiating attentiondeficit/hyperactivity disorders (ADHD) from other putative disorders, such as Sluggish Cognitive Tempo (SCT), in a large-scale adult sample (Barkley, 2012). Overall, the measures just cited are promising developments for the field, although careful analysis of item content is also needed, as ostensibly overlapping domains may nonetheless mask divergent association with other variables. For example, Weissman (2009) noted how

three “work impairment” scales all differentiated healthy from mentally ill adults but were only modestly intercorrelated themselves because they assessed different aspects of the work experience (e.g., workdays lost versus performance/problems at work versus interest/motivation).

Separating Impairment from Definitions of Mental Disorder Paralleling the struggle to develop comprehensive and valid systems of impairment, theorists and diagnosticians have grappled with the role of impairment in taxonomic systems for mental health. In 2009, a special section of the journal World Psychiatry was dedicated to discussion of a forceful proposal by Üstün and Kennedy (2009) for the DSM system to separate judgment of disability or impairment from the diagnosis of mental disorder, in a manner similar to the ICD and ICF systems. The authors noted that the DSM's inclusion of the “clinically significant impairment” criterion creates several inconsistencies with diagnosis in physical medicine, in addition to inconsistencies with the international systems used by the WHO. Reactions to this proposal were broadly supportive of this ideal: For example, Lehman (2009) noted that the current system of requiring clinical significance for a DSM diagnosis is analogous to leaving primary hypertension, absent other complications, untreated in physical medicine. However, respondents also raised various implementation concerns. More specifically, Narrow, Kuhl, and Regier (2009) noted that the ICF domains contain content very closely related to symptom criteria for several DSM disorders, making the separation of impairment and disorder difficult in practice. In commenting on the above-mentioned proposal, Wakefield (2009) pointed out that DSM implicitly allows the presence or absence of impairment to distinguish between disorder and nondisorder via not otherwise specified (NOS) categories (changed to other specified and unspecified disorders in DSM-5): that is, individuals can be diagnosed if impaired but not meeting full symptom criteria. Although this is desirable from the standpoint of maximizing access to diagnosis-dependent services, it potentially undermines the scientific basis of the DSM system. There is no simple solution to this problem, and Wakefield noted that even the ICD system contains many criteria that are impairment-related, despite the overarching goal to separate out impairment into the ICF. Ultimately, Wakefield argued for a close concept analysis of each disorder to determine how best to assess underlying biological/evolutionary dysfunction, as well as impairment.

Associations Between Psychopathology and Functional Impairment: Empirical Evidence Several recent reports have focused on the relations among symptoms of mental illness (both categorical DSM/ICD as well as broad dimensional problems) and measures of functional impairment. In general, these studies either demonstrate significant impairment from subthreshold conditions, replicating the seminal report by Angold, Costello, Farmer, Burns, and Erkanli (1999b), or else investigate what type or class of problems may result in more or less impairment. As an example of the former, Balázs and colleagues (2013) reported data

from the large-scale Saving and Empowering Young Lives in Europe (SEYLE) study on subthreshold depression and anxiety and functional impairment (assessed via the extended form of the Strengths and Difficulties Questionnaire). The authors found that both subthreshold anxiety and subthreshold depression were associated with increased risk of functional impairment controlling for the association across the two illness domains as well as demographic variables. Other work has demonstrated significant functional impairment even at the first visit for early-intervention services for affective disorders (Hamilton, Naismith, Scott, Purcell, & Hickie, 2011). In addition, the large-scale international Global Burden of Disease study began inclusion of CD and ADHD impacts in 2010, showing (despite various assessment and methodological challenges) that both conditions result in significant disability and qualityof-life impairments in youth, particularly males (Erskine et al., 2014). Other investigators have explored how the association of symptoms and impairment may differ across disorder categories. Often this work is motivated by relatively low to moderate associations between disorder symptoms and impairment, which can result in dramatic lowering of prevalence rates, depending on whether clinically significant impairment is included as part of the diagnostic criteria. For example, Lewandowski, Lovett, and Gordon (2009) gave a detailed discussion of impairment in ADHD, documenting that correlations between symptoms and impairment measures are generally no higher than .5 (i.e., 25% of shared variance) and are often substantially lower in this domain. Depending on how cutoffs for impairment are used, dramatic differences in ADHD rates will be seen (Gathje, Lewandowski, & Gordon, 2008). Low associations between symptoms and impairment have also been documented in the domain of separation anxiety disorder, with minimal incremental validity of impairment for future symptoms; the authors argued that distress, rather than impairment, may be a more focal predictor for this particular disorder (Foley et al., 2008). Relations among distress and impairment have been broadly studied by Ezpeleta, Reic, and Graner (2009), with results suggesting demographic differences in distress levels (with females and adolescents displaying greater distress overall) and disorder differences in impairment: broadly, externalizing disorders such as ADHD, ODD, CD, and substance use problems were associated with more significant impairment than most anxiety disorders but not more than major affective illness. In fact, the negative impact of ODD and CD (independently) on functional impairment across multiple domains has been corroborated longitudinally into young adulthood by Burke and colleagues (2014) using data from the Developmental Trends Study. Further complicating matters, associations among problem domain and impairment may vary depending on the specific measure of impairment used and the informant assessed. For example, Francis, Ebesutani, and Chorpita (2010) analyzed scores from the highly detailed Child and Adolescent Functional Assessment Scale (CAFAS) and the GAF in a diverse sample of youth ages 5 to 17 years living in Hawaii and referred for mental health evaluation. The study compared the proportion of youth meeting standard cutoffs for serious emotional disturbance on each measure for both broad externalizing and internalizing disorders. Of note, Francis and colleagues reported that the CAFAS identified more youth with externalizing disorders as meeting serious emotional disturbance criteria compared with internalizing

disorders, whereas the GAF showed more equivalent rates across broad disorder categories. Unfortunately, absent both additional information and a gold standard for serious disturbance, it is difficult to ascertain whether this result represents an oversensitivity of the CAFAS to externalizing problems versus an undersensitivity of the GAF to the same domain. Regarding multiple informants, research suggests that the informant specificity characterizing symptom ratings may also generally apply to impairment ratings, making for complex patterns of associations when both symptoms and impairment are assessed across multiple settings (Gadow, Kaat, & Lecavalier, 2013). Although work on impairment and symptoms to date has been largely cross-sectional, Cleverley, Bennett, and Duku (2013) reported a longitudinal study of functional impairment and internalizing symptoms through early adolescence (ages 10–15) in a large-scale, nationally representative and nonreferred Canadian sample. Measures of internalizing were available each year whereas impairment assessment was more limited (a single item for each of three domains: academics, peer relationships, and organized activities, with significant problems in any domain at any time resulting in classification of “impairment”). However, impairment status was associated with baseline internalizing symptoms, controlling for change in internalizing over time. Thus, there was consistent evidence for a main effect of impairment on internalizing (but not growth in internalizing) in this study. Other work examining associations between symptoms and impairment using epidemiological sampling has been conducted by Stringaris and Goodman (2013) in the United Kingdom, finding modest associations between symptoms and “impact” (impairment + distress) as well as incremental prediction of impact to future disorder status above symptom counts. Relatively little work has centered on either cultural variation in impairment or racial and ethnic differences in the relation between symptoms and impairment. When such efforts have been made, results suggest the importance of taking into account cultural context when discussing impairment. As a brief example, Jayawickreme, Jayawickreme, Atanasov, Goonasekera, and Foa (2012) demonstrated incremental validity in predicting WHODAS 2.0 functional impairment by using local idioms of distress in addition to Western-based symptom measures in an adult Sri Lankan sample. Additionally, data from an evaluation study of community mental health services for children and families in the United States suggested that relations between behavioral/emotional strengths and functional impairment may vary by ethnicity (Barksdale, Azur, & Daniels, 2010). There is evidence that youth social withdrawal may be less impairing in Asian cultures as compared to Western cultures and that parents' likelihood of seeking professional help for their children's symptoms (a potential indicator of impairment) may vary cross-culturally (Rapee et al., 2012). However, not all investigations show evidence for cross-cultural differences related to impairment, as Zwirs et al. (2011) found evidence for similar relations between symptoms and impairment (though mean differences of each) across ethnic Dutch and immigrant Moroccan, Turkish, and Surinamese children living in the Netherlands.

NIMH Research Domain Criteria Project Although it is distinct from the DSM and ICD systems, there has been a high-profile effort to

link mental disorder research with a system of constructs developed to emphasize multiple levels of analysis and, in particular, neurobiological substrates of disorder. The National Institute of Mental Health (NIMH) Research Domain Criteria (RDoC) project is a collaborative effort among researchers from basic science, clinical, and translational backgrounds to move diagnostic efforts closer toward identifying the pathophysiological mechanisms underlying disorder (Insel et al., 2010; Sanislow et al., 2010). This work is grounded in the long-known study of endophenotypes: well-specified physiological or behavioral indices that act as intermediaries between genetic influence and clinical disorder (Gottesman & Gould, 2003). The RDoC initiative aims to define dimensional constructs relevant to disorder at multiple levels of analysis, making the endophenotype concept more explicit and potentially more useful (Insel & Cuthbert, 2009). Levels of analysis as defined in the RDoC include genes, molecules, cells, circuits, physiology, behavior, self-reports, and paradigms: Each is crossed with multiple constructs within each broad domain to generate matrices of information to guide further research, construct validation, and ultimately new diagnostic classification systems (Cuthbert & Kozak, 2013). There are promising connections between the RDoC constructs chosen for initial study and constructs central to this chapter. The initial RDoC domains/systems include positive affect, negative affect, cognition, social processes, and regulatory systems (Sanislow et al., 2010). Several narrower constructs within these broad domains have clear connections to competence in our framework of success in developmental tasks. In particular, the RDoC domain of social processes, which includes affiliation and attachment, reception and production of facial and nonfacial communication, and perception and understanding of self and others (NIMH, 2014), contains multiple connections to our focus on social competence. However, the RDoC's emphasis on constructs closely linked to biology and neuroscience has tended to focus attention on narrower, behaviorally assessed traits/abilities such as behavioral inhibition and the recognition of facial expressions (see Lilienfeld, 2014, for related discussion of the limitations of RDoC's biological focus). Although the RDoC project is a sweeping and ambitious step toward increased validity of clinical syndromes, to date it has been approached in a largely nondevelopmental manner, and its ultimate utility may be gauged in part by how well it can be integrated with a comprehensive science of developmental psychopathology. Early steps in such integration have been taken by Franklin, Jamieson, Glenn, and Nock (2015), who highlight, among other points, the importance of multiple methods and integration of findings across levels of analysis.

Impairment: Conclusions As discussed in the second edition of this chapter (Masten et al., 2006), some of the debates surrounding the role of impairment in psychiatric diagnostic systems parallel those in the realm of competence. Issues of construct validation, choice of specific assessment measure, and aggregation of data from multiple informants are crucial factors in both areas. Certainly the field remains far from a consensus on any of these issues with regard to impairment. Indeed, discussions such as the debate over inclusion versus exclusion of impairment in DSM-5 diagnostic criteria suggest that research and theory on impairment remains in the fairly early

stages. However, the presence of both widely cited empirical demonstrations of the importance of impairment (e.g., Angold et al., 1999b) and compelling recent conceptual reviews (e.g., Rapee et al., 2012) offer promise for the future. More specifically, when one studies the actual domains and item content assessed in measures of functional impairment, the links with the developmental study of competence are obvious. Indeed, it is not much of an exaggeration to say that impairment is the negative pole of competence. However, we can go beyond this basic relation to compare and contrast the domains that are deemed salient by each broad research tradition. A complication here is the relatively nondevelopmental nature of much impairment research, with the notable exception of careful work creating age-based normative data for measures of adaptive behavior such as the VABS (Sparrow et al., 2005) and the Scales of Independent Behavior-Revised (SIB-R; Bruininks et al., 1996). The broad developmental tasks of academic functioning and social relationships (broad peer acceptance and/or close friendship) have prominent placement in nearly all youth impairment instruments. In turn, many impairment measures focus on domains of youth's lives, such as self-care, leisure activities, and broader community engagement, that are less often seen as central in other developmental task work, with one exception being the positive youth development framework (e.g., R. M. Lerner, von Eye, Lerner, & Lewin-Bizan, 2009), discussed in more detail in the “Conceptual Models” section of this chapter. Finally, impairment has clear relevance for interventions with disturbed youth, as it is conceptualized as the visible impact of underlying disorder, is often the reason for referral for services, and often is (or should be) the primary target for intervention (Barkley, 2009). In fact, a recent discussion of evidence-based interventions argues that the standards for what qualifies as evidence-based should include measurements of practically significant reductions in functional impairment in addition to clinically significant reductions in symptoms (Becker, Chorpita, & Daleiden, 2011). Of note, Becker and colleagues described the relative lack of an evidence base on reductions in impairment among classic randomized clinical trial (RCT) designs, which is due to both omissions of such measures from many study designs and greater challenges in influencing impairment for those studies that include such instruments. Others have noted that reducing impairment can be crucial for the “social validity” of an intervention (essentially, participant and family buy-in) and argue that measures of impairment should be included in RCT designs and should also be tailored in some degree to the particular aspects of functioning directly or indirectly targeted by the intervention (Fabiano & Pelham, 2009).

Cascade Models Conceptual and Statistical Background Cascade models are both conceptual models and research designs that emphasize spreading effects across domains, generally through inclusion of three or more target constructs at three or more time points in a single study (Masten & Cicchetti, 2010). These models have roots in theoretical conceptualizations of cumulating problems or advantages across major adaptive domains (Cicchetti & Schneider-Rosen, 1986). As in many other domains, methodology in this

area remains under close scrutiny and continued development: Conceptual and statistical discussions are prevalent and important for adequate communication and advancing theory. For example, Gonzalez (2009) has provided a thoughtful discussion of various methods for the analysis of transactional models: Like cascade models, transactional models emphasize interconnection of distinct constructs over time, although the two classes of models are not identical (Sameroff, 2009). In particular, Gonzalez highlighted the need to consider individual heterogeneity and nonlinearity in testing theories of development and change. Cascade models are statistically similar to longitudinal mediation models, especially the subset of models in which the predictor variable, mediator, and outcome are all measured on at least three occasions. In a series of methodological papers, David Cole and Scott Maxwell have provided comprehensive discussions of the major analytic issues related to longitudinal mediation models, highlighting both their clear advantages over models using only crosssectional data and the potential concerns and additional assumptions needed for accurate inference. To begin with, they noted that prior levels of the dependent variable represent an unmeasured confound in most cross-sectional analyses of mediation, which increases the challenge of assessing the magnitude of indirect effects in such designs (Cole & Maxwell, 2003). Subsequently, Maxwell and Cole (2007) demonstrated that estimates of the magnitude of the indirect effect from cross-sectional analyses can be either positively or negatively biased relative to the estimates from longitudinal data, depending on a number of factors, such as the stability of the predictor and the mediator. They also illustrated that the presence or absence of such bias can be substantial and difficult to predict in advance without making stringent assumptions. Cole and Maxwell (2003) went on to emphasize several additional points relevant for our discussion. First, they noted the importance of timing in longitudinal assessments, including the obvious but rarely acknowledged point that timing of measurement does not always equate with timing of causal action for a construct (see R. M. Lerner, Schwartz, & Phelps, 2009, for related discussion). Thus, whenever possible, researchers should engage in pilot work to establish the minimum estimated lag time for both a predictor's effect on a mediator to emerge and (separately) for the effect of the mediator on the outcome to emerge, and they should design longitudinal assessments to bridge those time intervals, which will not necessarily be equal. Longer and shorter intervals than these standards can bias estimates of indirect effects, related to the bias introduced by a cross-sectional design. We should point out that, in practice, statistically optimal lag times remain unclear in many areas of developmental psychopathology, a fact that parallels concerns about our lack of knowledge of causal structure and ambiguous constructs generally (e.g., Aalen, Røysland, Gran, & Ledergerber, 2012). Second, Cole and Maxwell also emphasize the importance of estimating and reporting overall indirect effects for studies designed with more than three time points, in which several mediating paths (i.e., unique tracings from the predictor to the outcome) are estimated; others have stressed the calculation of confidence intervals for total indirect effects in longitudinal designs (Little, Preacher, Selig, & Card, 2007). Third, they discuss the complex and unpredictable effects of measurement error, noting that mediators in particular should be assessed using multiple broad manifest indicators in a latent variable framework whenever possible and that informant

variance should be noted as a possible biasing factor. Building in part on this work, Selig and Preacher (2009) discussed the wider use of longitudinal mediation analysis in developmental research. In particular, they highlighted the fact that cross-lagged panel designs do not directly assess intraindividual change, for which latent growth curve models (Preacher, Wichman, MacCallum, & Briggs, 2008) are needed. Panel designs also do not allow change in one construct to serve as a predictor of another construct, for which growth curve models or latent difference score models (McArdle, 2009) are needed. Importantly, Selig and Preacher detailed how mediation models can be combined with either latent growth or latent difference models to answer developmental questions. Although we detail examples of some of these approaches in this chapter, these methods remain underused and may offer new windows for elucidating the nature of relations between competence and psychopathology from a perspective of individual differences in intraindividual change. Note that the relevant interpretive challenges do not simply go away with use of these methods, since each type of analysis allows for several different types of indirect effects (e.g., effects via slope versus intercept for growth curve models; effects via latent change scores versus initial values for latent difference models). It is important to stress that the cascade models discussed in this chapter do not all follow a single, prescriptive analytic framework. However, the majority of studies conducted within this conceptual area have a longitudinal panel design structure with at least three constructs of interest assessed on at least three occasions of measurement. In addition, the estimation of indirect effects, through various paths, is of substantial interest for this area of work. In reviewing the work to be discussed, there is clearly a greater need to incorporate some of the tenets of Cole and Maxwell's (2003; Maxwell & Cole, 2007) methodological studies to obtain the most accurate estimates of effects in the future. For example, even modelers who are most interested in testing whether effects from one domain assessed early in a longitudinal study (e.g., externalizing behavior problems) eventually spread out to affect multiple later domains are implicitly testing for the presence of a nonzero longitudinal indirect effect. However, given the current state of the science, we also applaud the use of multiple analytic methods in this domain. For example, some work uses latent growth curve modeling to test whether effects in one domain “spill over” to affect the change in a separate domain (Klimes-Dougan et al., 2010). Other work makes use of intervention designs to test whether hypothesized cascades can be interrupted (McClain et al., 2010; Patterson, Forgatch, & DeGarmo, 2010). Cole and Maxwell (2009) provided further elaboration of methodological issues related to studying longitudinal relations among constructs. Their work highlighted the sensitivity of results to variations in individual patterns of change and specified several additional potential analytic approaches (e.g., random coefficient modeling, latent state-trait modeling) that are of potential interest in this domain. Relatedly, Selig, Preacher, and Little (2012) emphasized the explicit modeling of varying time lags across individuals and potentially the random assignment of participants to different measurement lags, if one's focus is on elucidating the timing of a developmental process. The largely variable-focused nature of these methods can be complemented by person-oriented methods that aim to identify relatively homogeneous subgroups of a sample, either concurrently or via patterns of change over time (Masten, 2001;

Nagin & Tremblay, 1999; Von Eye & Bergman, 2003). Unfortunately, none of the statistical techniques just described eliminate well-known and fundamental causal inference problems for nonexperimental data. Social scientists have long been aware of the need for experimental, or quasi-experimental, manipulation of constructs for establishing causal inference, and new statistical and theoretical work has drawn closer attention to the conditions under which causal inference can be justified (Foster, 2010). Recent advances suggest ways in which careful selection of exogenous predictors in a model can move the field closer to causal inference, given specific assumptions (Pearl, 2000, 2012). This work on causal mediation analysis, largely developed outside of psychology, is nonetheless having an impact on the field. In a key contribution, Imai, Keele, and Tingley (2010) detailed the assumptions required to make causal attributions from both experimental and nonexperimental data and provided routines to conduct sensitivity analyses for key assumptions and bootstrapping to estimate the magnitude of mediating effects through a psychological research example. Few of the studies to be reviewed have been directly informed by these approaches. However, prevention scientists have considered this body of work in efforts to clarify mechanisms of mediation and reduce selection bias (Coffman, 2011; Crowley, Coffman, Feinberg, Greenberg, & Spoth, 2014), and a highly accessible treatment of modeling confounds in resilience research has been presented by Card and Barnett (2015). This is clearly an area for further research and application; nevertheless, we caution researchers in this area to be conservative in their discussions of causality and to detail all attempts made to justify the necessary analytic assumptions.

Empirical Findings Here we review findings from selected empirical investigations that are consistent with our understanding of testing developmental cascades. We highlight in particular the series of papers published in two special issues of Development and Psychopathology in 2010 (Masten & Cicchetti, 2010) as well as selected additional papers. Because these papers focus on a variety of constructs, samples, and age ranges, they are somewhat difficult to classify. However, given this chapter's focus on competence and psychopathology, we have organized this discussion by areas of competence that have been assessed. We first focus on studies including academic competence in patterns of relations with risk and psychopathology, followed by studies including social competence and studies testing both academic and social functioning. We then discuss other studies that have examined a wide range of target constructs. Where possible, we draw attention to convergent empirical evidence across age ranges as well as the implications of differences in statistical/data analytic methods. In most cases, research is focused on estimating cross-domain longitudinal paths between competence domains and either externalizing or internalizing problems. Testing Cascade Effects through Academic Competence Several investigations have examined the longitudinal links among behavioral/emotional problems and academic competence. Masten and colleagues (2005) reported results from the Project Competence longitudinal study, demonstrating that childhood externalizing problems

were negatively associated with later academic achievement in late adolescence/emerging adulthood, and academic competence in turn was negatively associated with internalizing problems in young adulthood. Key cross-domain paths remained statistically reliable when adding a subset of potential common-cause variables, namely socioeconomic status (SES), intellectual ability, and quality of received parenting. In addition, the authors noted a negative longitudinal path from internalizing to externalizing, one possible interpretation of which was the personality construct of behavioral inhibition simultaneously acting as a protective factor for externalizing problems but as a risk factor for internalizing problems (Kerr, Tremblay, Pagani, & Vitaro, 1997). Following Masten and colleagues (2005), in recent years the associations among externalizing, internalizing, and academic competence have continued to be studied across different samples and age ranges. A subset of these studies have extended this work by longitudinally testing the role of alcohol and substance use in cascade findings that relate to academic competence (or failure) and broader externalizing problems, given long-standing work on the associations across these domains. Other work has expanded the range of samples, age ranges, and assessment intervals that are employed. Finally, as discussed in subsection “studies including both academic and social competence,” some work has examined reciprocal and joint paths among academic competence, social competence, and behavioral and emotional problems. An investigation by Moilanen, Shaw, and Maxwell (2010) tested longitudinal links among externalizing problems, internalizing problems, and academic competence in a sample of lowincome boys and their families, with rich multi-informant assessments of each construct spanning ages 6 to 12. In the best-fitting structural equation model for their data, Moilanen and colleagues reported consistently high rank-order stability of each of their primary constructs across this developmental period. Above and beyond stability and concurrent association, the authors reported negative cross-domain paths from externalizing problems to academic competence (from ages 6–10) and from academic competence to internalizing problems (from ages 8–11), as well as “feedback” negative paths from academic competence back to later externalizing. Contrary to Masten and colleagues (2005), positive externalizing-to-internalizing paths were observed at the start and end of the study. However, consistent with Masten and colleagues' (2005) results, primary cross-domain paths were largely unchanged when including additional early constructs of neighborhood advantage, IQ, and parenting in the model. Moilanen and colleagues (2010) interpreted their results in terms of “adjustment erosion,” positing that early externalizing problems in particular could have both direct and indirect effects on functioning in the classroom, especially during the transition to formal schooling and the transition to middle school, transitions that bookended their assessments. Starting at a similar age range as Moilanen and colleagues (2010) but extending farther into adolescence, Englund and Siebenbruner (2012) reported cascade model results from the Minnesota Longitudinal Study of Risk and Adaptation, an intensive long-term study of youth at developmental risk due to poverty. Findings were again broadly consistent with prior work, especially in terms of academic failure (here, at age 9) predicting longitudinally to future externalizing and internalizing problems as well as a negative internalizing-externalizing link consistent with results of Masten and colleagues (2005); covariates included SES, minority

status, and age of mother at participant's birth. However, in the Englund and Siebenbruner (2012) study, longitudinal links from externalizing to academic competence were not observed, although the two constructs were associated at the first time point in their model (age 7; standardized path estimate of –.22). Importantly, and consistent with much prior work, Englund and Siebenbruner reported that prediction of substance use in adolescence was mediated through earlier (and concurrent) externalizing problems, although they also noted an unexpected positive association from academic competence to use versus nonuse of alcohol, suggesting that presence of alcohol use, distinct from amount of drinking, may have been quasinormative for this at-risk sample. Other work examining links between academic achievement and substance use, although not conducted from a full cascade framework, has supported a consistent longitudinal association between marijuana use and academic underperformance throughout high school in an affluent sample (Ansary & Luthar, 2009). Other work in this area has focused more closely on the longitudinal associations among academic competence and either broad internalizing problems or key subsets of internalizing, such as depressive symptoms. For example, Hishinuma, Chang, McArdle, and Hamagami (2012) examined the longitudinal links between academic functioning and depressive symptoms in a large (N > 7,000) sample of Hawaiian youth across grades 9 to 12. In contrast to the traditional cross-lagged structural equation modeling (SEM) design, this study employed a sophisticated bivariate latent difference score model, although it should be stressed that this analytic approach does not fully remove the potential criticism of unobserved confounders preventing causal inference. Of interest, Hishinuma and colleagues (2012) found stronger evidence for depressive symptoms being associated with future changes in academics (with somewhat stronger effects for native Hawaiian youth), which contrasts with several of the cross-lagged studies already discussed. It is unclear whether this discrepancy in findings may be related to analytic differences or perhaps differences in the sample or assessment: For example, the Hishinuma et al. study relied primarily on self-report measures for both academics and depressive symptoms. In addition, latent difference scores were set up annually; as discussed by Cole and Maxwell (2003), lag time between assessments may have influenced the pattern of findings. More frequent assessments entered into an extensive analysis of data from the Pittsburgh Youth Study (a longitudinal study of over 1,500 boys oversampled for antisocial behavior) from ages 11 to 15 (Defoe, Farrington, & Loeber, 2013). The authors tested a series of models linking SES, academic achievement, hyperactivity, delinquency, and depressive symptoms annually over this age period. Although due to complexity some constructs were split into separate models, the results showed impressive consistency in terms of direction of effects. Briefly, evidence was obtained for a longitudinal chain of effects starting with early joint influences of hyperactivity and low SES on academic achievement to delinquency and finally to depressive symptoms. Although cross-domain paths tended to be small in magnitude, they were obtained in excess of both single-year and 2-year lagged continuity estimates, which themselves were generally moderate in strength. Other work has examined academic achievement itself as an outcome of cascading effects that cumulate from early childhood. Notable for its prediction from very early in childhood, a

report by Bornstein, Hahn, and Wolke (2013) demonstrated indirect effects from habituation efficiency at age 4 months to cognitive testing at 18 months and 8 years (indirectly affecting and affected by 36-month behavior problems) to academic achievement at age 14. In this study, using data from the Avon Longitudinal Study of Parents and Children, behavior problems and maternal education had both direct and indirect effects on later academic achievement. Early cognitive measures had small but reliable effects on childhood intelligence, which was the largest single predictor of academics in adolescence. Further work by this same research team has demonstrated two additional findings of interest. First, infants' early physical development and motor exploration (at 5 months of age) appear to act as catalysts for their later intellectual development, with cascading effects seen to IQ at ages 4 and 10 years and subsequently to academic achievement at age 14 (Bornstein, Hahn, & Suwalsky, 2013c). Second, language abilities in early childhood showed cross-domain paths to subsequent internalizing problems in two samples controlling for a series of potential common-cause variables (Bornstein, Hahn, & Suwalsky, 2013b). Although language facility has only rarely been studied from the cascade modeling perspective, it is a promising candidate for future research due to its status as a key developmental task of the early childhood years (McCormick et al., 2011) and is a clear foundational task for academic achievement. In addition, longitudinal evidence using bivariate growth curve models has shown stronger prediction from language skills—specifically, receptive vocabulary—to (reduced) growth of externalizing problems and inattention/hyperactivity from ages 7 to 13 (Petersen et al., 2013). Summary of Academic Competence Cascade Evidence To date, several independent data sets have shown evidence for cross-domain longitudinal prediction from externalizing behavior problems to lowered academic competence. In addition, there is generally consistent evidence for links from lowered academic competence to increased internalizing problems, although this association is less consistent than the former one. Effects appear to be stronger when indexed with somewhat longer time intervals, excepting very young childhood; indeed, in this body of work, the chosen time intervals are not always justified by developmental considerations, and an argument can be made to vary assessment intervals based on the expected pace of change in a given developmental period. Overall, however, this work affirms the role of adequate academic performance as a key developmental task in the overwhelmingly North American and Western European societies from which these data have been drawn. In the majority of cases where gender differences in cross-domain paths have been tested, evidence has not been found for differences. However, it should be emphasized that power analysis for SEM is evolving, and it is not entirely clear how large sample sizes need to be to convincingly rule out Type II error for tests of gender differences. Target sample size will depend on the predicted effect size of key model paths, which is an extension to this research that is still in early stages. Not all investigations have found robust associations between early externalizing problems and academic achievement: Duncan and colleagues (2007) pooled results across six data sets and found negligible predictive power of early socioemotional variables (including internalizing and externalizing) to academic achievement later in elementary school when also controlling

for early school readiness variables, attention, and a wide set of demographic control variables. Rather, preacademic skills and attention skills appeared to be stronger predictors. It is likely that choice of covariates and potentially the direct versus indirect assessment of socioemotional variables play a role in differences in findings in this domain, as discussed further later. Although not focused on broad internalizing problems, a separate meta-analysis by Huang (2015) showed evidence for a modest negative relation (mean r = –.15) between academic competence and subsequent depression, which was reduced when including prior depression in the analysis and not moderated by gender, ethnicity, birth cohort, or measurement strategy. Testing Cascade Effects through Social Competence Cascading relations among social competence and behavioral/emotional problems have been even more frequently studied than links with academics. Using the same sample and parallel analytic techniques to the Masten et al. (2005) study, subsequent Project Competence research examined the longitudinal links among externalizing, internalizing, and social competence (Burt, Obradović, Long, & Masten, 2008), finding evidence for moderate to strong negative longitudinal associations from social competence to internalizing problems, both from childhood to adolescence and from emerging adulthood to young adulthood. Although longitudinal links were not seen between externalizing problems and social competence in this study, the latter constructs were strongly negatively associated (standardized coefficient of –.48) in the initial middle-childhood assessment. A limitation of the study by Burt and colleagues (2008) was reliance on relatively large time lags between assessments, on the order of several years. In addition, although constructs were assessed comprehensively through multi-informant and multimethod assessments, the study took a big-picture view of key constructs, treating social competence as a single measure (consistent with confirmatory factor analytic evidence for those data). However, other studies have examined related constructs, such as popularity and peer social preference, in more detail and across narrower time ranges. Generally speaking, this work confirms that failures in social competence in both childhood and adolescence can have negative links to later internalizing (and sometimes externalizing) problems. In addition, multiple studies have found longitudinal links from early externalizing problems to subsequent problems with peers. In each case, peer rejection and/or victimization are likely to be key elements. Following the core analytic strategy of Burt and colleagues (2008) but using annual assessments including family adversity exposure in the oldest cohort (ages 9–14) of the Pittsburgh Girls Study (a group oversampled for neighborhood disadvantage and followed annually), Obradović and Hipwell (2010) replicated an externalizing-social competenceinternalizing cascading effect across ages 9 to 12. In addition, child behavior problems predicted later family adversity exposure (e.g., parental depression, harsh parenting, and perceived parental stress) in these models, consistent with a “child effects” framework and with work to be discussed suggesting that preventing child problems is associated with changes in parental and family functioning (Patterson et al., 2010). Finally, pubertal status had negative links with internalizing problems early in the study, but the direction of this effect

shifted as the majority of the sample reached puberty, suggesting that pubertal effects may vary depending on the larger peer context. Similar to the Project Competence report, this study assessed a broad social competence construct made up of both peer and friendship social status and social self-worth, although competence assessment was solely self-report. Other work from a cascades framework has found evidence for social competence as a key predictor in middle childhood and has extended this work to examine diverse samples and potential moderators. For example, Rogosch, Oshri, and Cicchetti (2010) tested social competence cascades in a sample of maltreated and nonmaltreated but demographically matched youth. They found robust main effects of maltreatment on all major constructs (externalizing, internalizing, and social competence) at the first time point (ages 7–9) and evidence for social competence at ages 10 to 12 as a negative predictor of later externalizing problems; earlier internalizing was a negative predictor of social competence. Rogosch and colleagues (2010) were interested in predicting cannabis abuse and dependence in late adolescence, and showed both direct and indirect (through internalizing and externalizing) links from maltreatment history. In a separate study, Bukowski, Laursen, and Hoza (2010) demonstrated that friendship status moderated the effect of peer rejection on growth curves of peer-reported sad affect across 3 years of middle childhood. That is, among youth with high initial scores of peer avoidance and peer exclusion, depressed affect increased over time for youth without close friends but decreased for youth with at least one close friend. Although the latter study used a different analytic strategy from most studies discussed in this section, the use of growth curve models may be of special importance for future study of developmental cascades, and we consider several additional examples of this analytic strategy later. Taken together, the previous studies show the importance of social competence as an organizing construct in middle childhood across diverse samples and methods. Several investigations have pushed back the developmental time frame of cascade models, assessing social competence and peer relations earlier in childhood while sometimes still maintaining assessments through adolescence and into adulthood. This extensive time frame is particularly challenging since typical longitudinal structural equation models require similar measurement of constructs through the duration of the study, to avoid confounding change in measurement with true construct differences. In a study extending this broad work to the early childhood years (beginning at age 2), Blandon, Calkins, Grimm, Keane, and O'Brien (2010) examined links between early externalizing problems, developing emotion regulation and social skills (beginning at age 4), and later sociometric status. Consistent with greater fluidity of development and less differentiation in the early years, Blandon and colleagues found evidence for less stability of constructs than similar studies with older participants and noted very high associations between early emotion regulation and social skills measures, both of which serve to limit detection of any nonzero cross-domain effects. Nonetheless, negative cross-domain paths were noted from social skills to later externalizing problems at multiple time points, providing evidence for the importance of these emerging skills early in childhood. Two additional studies have examined social competence and peer relations in early childhood and their associations with later behavioral/emotional problems. First, Bornstein, Hahn, and Haynes (2010) showed evidence for negative links from age 4 social competence to age 10

and 14 externalizing and internalizing problems, effects above the moderate stability in each domain and which were unchanged when including child intellectual functioning, maternal education, and maternal social desirability in the path model. The authors noted that their results reinforced the historical importance granted to early social competence by attachment researchers (e.g., Sroufe, Egeland, & Kreutzer, 1990). The same research group demonstrated cascading effects from externalizing problems at age 4 to broad adaptive functioning (communication, socialization, motor and living skills on the VABS) at age 10, controlling for IQ (Bornstein, Hahn, & Suwalsky, 2013a). Second, Lansford, Malone, Dodge, Pettit, and Bates (2010) tested a series of paths connecting social information processing (SIP), peer rejection, and aggression over 12 assessment periods spanning kindergarten through third grade in the Child Development Project, building on earlier work from that sample and others examining SIP as a partial mediator of effects of early rejection on aggression (Dodge et al., 2003). In brief, Lansford and colleagues found evidence for a series of “snowballing” effects in which SIP deficits were associated with later rejection, which was associated with both aggression and further SIP changes, with aggression also providing a separate link to subsequent peer rejection. The authors noted that the lower stability of SIP relative to rejection and aggression implies that SIP should be a key target for prevention and intervention efforts. As evidenced by Lansford and colleagues (2010), closer waves of assessment allow for the development of feedback loops over time among constructs. In this vein, two further studies specifically highlight the “vicious cycles” that can arise during multiple developmental periods when child behavior problems are first linked with peer problems and those further exacerbate children's problems over time. Focusing on the early school years, van Lier and Koot (2010) aimed to clarify how different aspects of broad social functioning (social preference, victimization, and friendedness) related to externalizing and internalizing in a Dutch sample. Results from this study supported transactional links between externalizing and poor social preference over time, with initial externalizing predicting all three social constructs in expected directions. Victimization also became “entangled” with poor social preference in this study, suggesting complex relations among social competence constructs that also mediated longitudinal links between externalizing and internalizing problems, with key results invariant by gender. Follow-up work with this sample has replicated the bidirectional links between peer rejection and externalizing problems when including best friend's externalizing in the model, which did not add to cross-domain prediction at this age range (Sturaro, van Lier, Cuijpers, & Koot, 2011). Additional work by this team in a separate sample using linked growth curve models has shown that kindergarten conduct problems are linked to earlyelementary-school depressive symptoms via experiences of peer rejection (Gooren, van Lier, Stegge, Terwogt, & Koot, 2011). Relatedly, but moving into middle childhood and adolescence, an analysis of the Multimodal Treatment of ADHD (MTA) sample by MurrayClose and colleagues (2010) showed evidence for transactional effects between low social skills and aggressive behavior problems throughout the course of their study (ages 10–16). Aggression was also linked to antecedent inflated self-beliefs and prospectively to peer rejection in this sample, with paths invariant by gender and ADHD status. Finally, we wish to draw particular attention to two studies conducted by Dodge and

colleagues, using parallel analytic methods to test a long-ranging cascade of effects from early problem behavior to the development of serious violence and substance use problems in adolescence across two samples. This work is an example of how the cascade metaphor links to programmatic developmental work without necessarily prescribing a single analytic method. First, Dodge, Greenberg, Malone, and the Conduct Problems Prevention Group (2008) used this framework to examine a progression of effects from early contextual disadvantage to adolescent serious violence in the high-risk control sample of the well-known Fast Track intervention study. Key constructs spanned time periods from kindergarten through eleventh grade and were assessed via multiple informants and methods. Rather than traditional SEM, Dodge and colleagues chose a two-stage partial least squares (PLS) predictive modeling procedure, which seeks to maximize predictive power and has somewhat less restrictive assumptions as compared with other uses of confirmatory factor analysis and SEM. When matched with specific theory, PLS can provide important model testing, albeit with a different focus from typical SEM: one that emphasizes an aggregate index of items for prediction that is not hypothesized to be caused by a single latent variable. Results from this study suggested that each step of the hypothesized cascade—adverse context, early harsh parenting, poor school readiness, early school conduct problems, grade school academic failure, low parental monitoring in later grade school, affiliation with deviant peers in early adolescence, and violent behavior in adolescence—was related to the prior and subsequent steps. More specifically, each domain partially mediated the effects of the immediately preceding domain, and each added incrementally to the overall prediction of adolescent violence. Although most paths were invariant by gender, there was some evidence that social context and early parenting had greater prediction of later violence for females whereas early externalizing and school problems predicted for males. Among other contributions, these results confirm previous work on the bidirectional relations between school failure and externalizing behavior problems and provide compelling evidence for the early contextual and familial difficulties that precede the rapidly developing school problems for some externalizing youth. As the authors noted, a detailed cascading chain of effects also points toward places and time points that may be ripe for preventive intervention efforts. In fact, subsequent mediation work using the Fast Track intervention sample has shown that over 25% of the impact of the intervention on reducing adolescent antisocial behavior could be accounted for via three earlier socialcognitive processes: reducing hostile attributions, increasing competent responses to social challenges, and reducing favorability of aggressive responses (Dodge, Godwin, & the Conduct Problems Prevention Research Group, 2013). Second, a monograph by Dodge and colleagues (2009) employed parallel methods to study the development of illicit drug use in the previously mentioned Child Development Project sample, with associated commentary by Schulenberg and Maslowsky (2009). In this comprehensive study, Dodge and colleagues demonstrated how early child, family, and peer factors systematically build to predict the initiation of illicit substance use by the end of high school. Although effects at each time point were small to moderate in magnitude, the cumulative effect of these cascading influences is striking, with person-oriented analyses describing risk of substance use initiation that varied from 25% among youth following the lowest-risk sequence to 91% among youth following the highest-risk sequence. These analyses

were complemented by variable-oriented tests of the full model (i.e., from parenting problems in early childhood to adolescent substance use initiation via early child behavior problems, early peer problems, adolescent parenting problems, and association with deviant peers) again using composite variates and PLS modeling. Above and beyond the key cascading path of effects, continuity was seen for parenting problems and peer problems from childhood to adolescence. Similar to several of the studies reported earlier, mean gender differences were observed on many constructs, but developmental progression did not vary by gender. Perhaps most important, each piece of the hypothesized progression mediated (either partially or fully) the prior step in the sequence. Commentary on this study emphasizes the strong theoretical basis of the proposed sequence of effects, the utility of a parsimonious model with a low risk of overfitting to a specific data set, and the need for close study of developmental transitions (Schulenberg & Maslowsky, 2009). Summary of Social Competence Cascade Evidence An impressive body of work has arisen in recent years testing cascading effects between social competence and both externalizing and internalizing problems, spanning childhood and adolescence and employing several different methodologies. It is clear that, like academics, social competence represents a core developmental task beginning in young childhood and continuing throughout development. Several studies show the pernicious effects of externalizing problems on social adjustment, and multiple investigations show paths consistent with failure model theories that predict increased internalizing, and sometimes externalizing, problems from lower social competence. More specifically, findings from the social domain generally show somewhat less longitudinal stability in constructs but somewhat more cross-domain links as compared with academics. In addition, findings vary, albeit often in predictable ways, depending on the domain of social adjustment that is assessed (e.g., broad judgments of competence, peer rejection, social cognition). Of particular note are findings of “vicious cycle” links across time between peer relations and externalizing problems, and the programmatic work by Dishion, Dodge, and others that demonstrates how deviant peer association and constructs such as deviancy training and coercive joining of an antisocial peer group operate to reinforce and amplify aggression and antisocial behavior (Van Ryzin & Dishion, 2013). It should also be noted that the broad social competence construct has been linked to adult psychopathology in different ways. For example, although not tied to social competence as measured by many of the studies already discussed, Lenzenweger (2010) has proposed a theoretical cascade that posits absence of parental “proximal process” (reciprocal interactions with a child that foster psychological functioning) impairing the broad affiliative system observable via child temperament and leading to the development of schizoid personality disorder in adulthood. Studies Including Both Academic and Social Competence There are multiple reasons to highlight longitudinal cascade studies that include both social and academic competence domains: Longitudinal SEM studies statistically control for other

constructs in the model, thus providing more stringent tests of prior findings with single domains; relations between academics and social functioning themselves can be modeled in these studies; and key theories, such as the dual failure model for depressive symptoms, that rely on joint effects from failure in these two broad domains are more fully tested. In general, this research has supported dual failure theories in which the combination of problems with academics and one's peers interact to elevate risk for future mental health problems (e.g., Capaldi, 1992). Two reports from the Project Competence data set have explored the potential joint relations of academic and social competence with externalizing problems, internalizing problems, and the subsequent competence domain of effective work functioning. First, Obradović and colleagues (2010) demonstrated that earlier longitudinal cross-domain effects from academic and social competence in childhood and emerging adulthood to internalizing problems in young adulthood maintained when both domains were included in the statistical model, with academics also showing a positive longitudinal relationship to social competence. Second, subsequent work with this sample focused on the crucial young adult outcome of work functioning, demonstrating a longitudinal cross-domain path from childhood social competence to the domain of employment functioning (e.g., reliability, achievement) in emerging adulthood, which in turn was highly stable into young adulthood (Masten, Desjardins, McCormick, Kuo, & Long, 2010). Although academic competence did not show unique longitudinal prediction to the work domain, this may have been in part due to the high concurrent association (standardized path estimate of .59) in emerging adulthood. Other work has extended the dual failure cascade modeling to independent samples, while also looking at the timing of effects at narrower intervals. Burt and Roisman (2010) tested similar effects in the National Institute of Child Health and Human Development (NICHD) Study of Child Care and Youth Development (SECCYD), using maternal reports of behavior problems and social skills and standardized academic achievement testing at five time points from age 54 months to 15 years. They found evidence for indirect effects of early externalizing problems through academic competence/failure to internalizing in early childhood as well as indirect effects through social competence later in childhood (feeding back to externalizing and academic problems in adolescence), consistent with other work emphasizing social relations in predicting trajectories of externalizing and internalizing problems in the same data set (Fanti & Henrich, 2010). However, effects seen by Burt and Roisman were generally of small magnitude. At first glance, these findings conflict with Duncan and colleagues' (2007) report that also included the NICHD SECCYD as part of six meta-analyzed data sets. However, closer inspection of the analytic plan from each report suggests that it is the presence of specific covariates that affected whether the externalizing-academics association is present: In the Duncan et al. (2007) paper, earlier externalizing behavior (age 3) was also controlled for when examining age 4.5 aggression to kindergarten academic skills in this data set. Consistent with this finding, the early externalizing to academics link reported by Burt and Roisman dropped to marginal significance when the covariates of maternal sensitivity, income/needs ratio, and early cognitive ability were also included in the model. Related work from the Québec Longitudinal Study of Child Development, focused more

narrowly on the early school years, showed evidence for cascading effects from externalizing problems at age 6 through both academic problems and peer victimization at age 7 to internalizing problems at age 8. Peer victimization also fed back to externalizing, and early school readiness measures predicted negatively to internalizing 1 year later; again, crossdomain effect sizes tended to be small in the face of strong continuity and contemporaneous associations among constructs (van Lier et al., 2012). Subsequent work by this research team provides evidence for a cascade from peer rejection through (low) social self-concept to internalizing problems across second grade into third grade (Spilt, van Lier, Leflot, Onghena, & Colpin, 2014). Extending this work to slightly older youth, Vaillancourt and colleagues (2013) presented cascade evidence through the middle school years for internalizing problems leading to increased future peer victimization and for academic problems predicting future externalizing, both cascade effects persisting when controlling for household income and parental education. Finally, Chen and colleagues (Chen, Huang, Chang, Wang, & Li, 2010; Chen, Huang, Wang, & Chang, 2012) have presented two studies from a Chinese sample of elementary school students investigating these topics. Their first report supported paths from early aggressive behavior problems to lowered social and academic competence, and cross-domain longitudinal paths between social competence and academics, even in the context of high rank-order stability (Chen et al., 2010). A follow-up report supported a consistent link between aggression and poor peer relationships (seen at each annual window from grades 3–6) but also demonstrated that starting levels of aggression are associated with the growth of peer problems over the 4year duration of data collection (Chen et al., 2012). For all of these studies, although mean gender differences were frequently observed for key constructs, the longitudinal cross-domain paths generally did not vary by gender, suggesting the operation of similar processes. The authors highlighted the strict socialization away from externalizing problems in Asian culture (i.e., self-control and maintaining group harmony are highly valued) while noting their results were generally consistent with data from Western societies. Given the increased salience of peer relationships in adolescence, several longitudinal studies have focused on the preteen, teenage, and young adult years in testing links among behavioral and emotional problems and academic and social competence, the latter broadly defined to include deviant peer affiliation and peer victimization. First, work from the University of Oregon has modeled the longitudinal development of violent behavior in late adolescence via early problem behavior and academic failure, leading to gang involvement and deviancy training—the highly stable mutual reinforcement of deviant talk in friendships (Dishion, Véronneau, & Myers, 2010)—with consistent patterns of associations (despite mean differences) across African American and European American youth. Second, a major study of children at risk for alcohol problems using cascade methods has shown that effects of parental alcoholism on offspring substance use disorders in adulthood is mediated through adolescent substance use, academic achievement, and affiliation with a substance-using peer group (Haller, Handley, Chassin, & Bountress, 2010), with some differences in effects for alcohol versus other drugs (e.g., less evidence for selection effects for binge drinking as compared to illegal drug use). Third, data from the Seattle Social Development Project have shown

longitudinal prediction from depressive symptoms at age 10, through school failure in adolescence, to risk for a major depressive episode at age 21 (McCarty et al., 2008); however, this longitudinal mediated path was evident only for girls, in contrast to several other reports already noted. Although early social problems were also tested in this model and showed a zero-order correlation with earlier depressive symptoms, they did not show independent prediction to adult depressive episode risk for either gender. McCarty and colleagues (2008) noted that in their study, stronger prediction may have been seen for girls for both academic failure and early conduct problems to later major depression due to gender differences in base rates of the former problems: that is, girls identified with these problems may have been relatively more severe cases. Evaluation of one's competence in the school context may also have different consequences by gender on average, although other cascade studies have generally not shown supportive evidence for this assertion. Although not employing the typical cascade methodology, Nocentini, Calamai, and Menesini (2012) studied a bivariate growth curve model of delinquent behavior and depressive behavior across 3 years of Italian high school (grades 9–11). Interestingly, the authors found a somewhat different pattern of time-invariant versus time-varying effects for the two domains, with stronger evidence for distal effects (i.e., present at the start of the study and operating uniformly throughout) of academic and social failure on delinquent behavior but proximal effects (varying year to year, and sometimes operating indirectly) of dual failure for both delinquent behavior and depressive symptoms. Examining developmental task domains more broadly, Staff and colleagues (2010) emphasized the nature of role transitions across adolescence and young adulthood in analyzing data from the Monitoring the Future study, focusing on prediction of substance use. Similarly to comparing social selection versus social causation effects (Martin et al., 2010), a focus was on comparing role selection versus socialization explanations for associations between adult family and work roles and reduced substance use. Generally, their results showed strong support for the socialization hypothesis, showing that changes in social roles lead to changes in substance use controlling for several prior risk factors. Looking across domains, Staff and colleagues found that changes in family roles (e.g., marriage, parenthood) appeared to have stronger effects than did changes in social and work roles, controlling for age to account for differential timing of role transitions. Changes in socializing (spending evenings out for fun and recreation) and religiosity partially mediated these effects. This work suggests that the emerging adulthood years can be a key point of change for the continuation versus reduction of substance use, one aspect of their importance in developmental psychopathology more generally (Schulenberg, Sameroff, & Cicchetti, 2004). Summary and Future Directions: Academic and Social Competence Cascades The research reviewed is compelling in its broad support of dual failure models across wide ranges of the childhood and adolescent years. In particular, the inclusion of both academic functioning and social competence (or related constructs, such as peer rejection and victimization) in single longitudinal models demonstrates that these two aspects of global competence can themselves interact and that cross-domain effects seen via one domain are not

typically proxies for the other but rather may compound in interesting ways. For example, there is evidence that high academically achieving students generally self-select into peer groups with other high-achieving students by early adolescence (Véronneau, Vitaro, Brendgen, Dishion, & Tremblay, 2010). One implication from these findings, highlighted by van Lier and Koot (2010), is that early prevention efforts should be focused on enhancing both academic success and effectiveness in social relationships wherever possible, especially since peer affiliative groups may be easier targets for intervention compared to global peer status (Véronneau et al., 2010). Common to many studies of cascade modeling, important future directions for research include better delineation of the timing of effects, testing of moderators, including but not restricted to gender, and consideration of additional mediators. As an example of potential mediators, one study has shown evidence for increased parent-child conflict as a mediator of longitudinal externalizing-internalizing links from grades 4 to 8 in a model that also included academic and social competence (Yong, Fleming, McCarty, & Catalano, 2014). Regarding timing of effects, it is broadly seen that longer time intervals have resulted in stronger magnitudes of crossdomain predictive paths in this research. However, very few studies have explicitly planned time intervals to maximize strength of effect, as recommended by Cole and Maxwell (2003). It is possible that careful reanalysis of data sets with many time points, systematically testing subsets of varying time lags, could prove useful in clarifying this issue. However, any such work should be theoretically informed and should account for issues of multiple statistical comparisons rather than being purely data-driven. Future studies would also do well to heed Dodge and colleagues' (2008, 2009) method of systematically isolating and combining the most theoretically relevant predictive variables at each assessment period. In addition, it is an obvious but important point that the influence of a specific process cannot be tested unless it is accurately assessed. Thus, work by Dishion et al. (2010) demonstrating the key impact of deviancy training in close friendships (and highlighting the importance of identifying and preventing the growth of deviant peer clusters) was possible only by employing observational recording of adolescent dyadic friendships. Many of the studies reviewed are more limited in their range of measurement, often relying on primarily questionnaire-based assessment due to the resource cost of assessing multiple constructs across several time points (cf. Cox, MillsKoonce, Propper, & Gariépy, 2010). Additional Studies of Cascade Methodology: Links Across Generations and Diverse Constructs Although not focused explicitly on broad developmental task domains, several additional recent studies have advanced the literature on developmental cascades by extending aspects of the studies reviewed by use of new populations, new methods, and broader key constructs. Some of the ways in which cascade studies have been extended include studies of continuity and change in behavioral/emotional problems across generations as well as expanding established cross-domain links to encompass additional constructs such as life stress, effortful control (EC), and broad self-regulation. In addition, other studies represent potential methodological strengths, such as the use of latent growth curve models in studying cross-

domain effects and/or the inclusion of intervention or prevention designs. As a key example of studying intergenerational effects from this perspective, Garber and Cole (2010) presented longitudinal data from adolescents ages 12 to 18 years and their mothers, oversampled for a history of depressive symptoms. In a similar fashion to the study by Nocentini and colleagues (2012) described earlier, Garber and Cole distinguished between distal effects of maternal depression, which affected the intercept of adolescent depressive symptoms in a growth curve model, and proximal effects via other intermediate variables, which affected the slope of adolescent depressive symptoms in a growth curve model. Mediating variables included risk factors of negative family life events, overall family climate, and self-worth, all assessed in early adolescence. Importantly, the direction of effects was consistently from the risk factors to growth in depressive symptoms rather than the reverse. Despite the strengths of this design, one limitation is the inability to parse the interplay of genetic and environmental influences across generations. However, other work focused earlier in development (ages 9–27 months) used an adoption design to separate out risk for toddler externalizing problems from both biological maternal depressive symptoms (which appeared to operate via pregnancy risk) and risk from each adoptive parent's depressive symptoms (Pemberton et al., 2010). This work highlights the potential for bidirectional influences between children and their environment, which is consistent with work examining links between SES and adolescent problem behavior across two generations in the Iowa Family Transitions Project (Martin et al., 2010). Although not focused on biological variables, results from this model suggested support for both social causation (parent SES influencing adolescent behavior problems) and social selection (adolescent behavior problems influencing their adult SES) processes. Other research on intergenerational associations in antisocial behavior has emphasized potentially differing mediating pathways depending on which parent is the focal point of investigation. For example, Thornberry, Freeman-Gallant, and Lovegrove (2009a,b) demonstrated that in the ethnically diverse Rochester Intergenerational Study, parental delinquency and drug use in adolescence were associated with offspring externalizing behaviors primarily via effective parenting for mothers (nearly all of whom lived with their children) but via multiple mediating paths (parenting stress, effective parenting, and parent's age at first birth) for fathers who had high levels of contact with children; some influence of parental depressive symptoms was also noted. Importantly, the intergenerational associations were not present for fathers who had low contact with their children. Related intergenerational research from the Seattle Social Development Project has reported that although harsh discipline was associated with childhood externalizing problems across two generations, the intergenerational association of externalizing was mediated more by parental substance use than by harsh discipline (Bailey, Hill, Oesterle, & Hawkins, 2006, 2009). Despite some differences in findings, both data sets support the clear disruptive effects of parental substance use problems on offspring adjustment. Given the importance of behavioral and emotional regulation for the development of broad competence across childhood, it is not surprising that multiple investigations have examined longitudinal links between EC or other aspects of regulation and preceding, concurrent, and

subsequent externalizing and internalizing problems. Examples from this literature include a study of 18- to 42-month-olds by Eisenberg and colleagues (2010) that examined parenting, EC, and internalizing and externalizing problems at three time points. Briefly, study results showed support for maternal unsupportive reactions during an observational session negatively predicting subsequent EC; however, a mediating role of EC was not evident. This may have been due to high rank-order stability among the tested constructs in this sample (e.g., many standardized paths >.80). Work from the same research team has focused on later childhood, using bivariate semiparametric mixture modeling techniques to examine subgroups of youth with distinct trajectories of externalizing and two indices of EC: attention focusing and behavioral persistence (Zhou et al., 2007). This work suggests that although the population association of EC and externalizing may be driven by a subgroup of well-adjusted youth who are persistently high on the former and low on the latter, trajectories of externalizing for youth with more mixed longitudinal profiles of EC are more variable. Other work supports the importance of behavior regulation by demonstrating mediating effects of self-regulation in later years on associations between cumulative risk and externalizing, controlling for earlier externalizing problems (Doan, Fuller-Rowell, & Evans, 2012). In addition, work from a positive youth development (PYD) perspective analyzing the well-known 4-H study of Positive Youth Development has shown cross-domain longitudinal links across late elementary and middle school from positive parenting, to child internal self-regulation, to an integrative PYD construct (Lewin-Bizan, Bowers, & Lerner, 2010), which we return to in more detail when discussing prevention. Cascade research has also examined how individual- and family-based stressful life events interact with externalizing and internalizing problems over time. Yates, Obradović, and Egeland (2010) presented data from the Minnesota Study of Risk and Adaptation, drawing on three assessments spanning 24 to 72 months of age, showing both direct and indirect (through child dysregulation) longitudinal paths between maternal contextual strain (a combination of stressful life events, social support, and romantic relationship quality) and child externalizing. In addition, Herrenkohl and colleagues (2010) used data from the Seattle Social Development Project to demonstrate independent prediction from family adversity and conduct problems in middle childhood to health risks and service use in young adulthood. Focusing on self-reported stressful events, Brody, Chen, and Kogan (2010) demonstrated that self-reported negative emotions mediated the association between life stress, including racial discrimination, and the slope of a risk behavior growth factor (alcohol use, drug use, and having sex under the influence) in a sample of rural African American youth assessed through late adolescence and early adulthood: Data were collected as part of a family-based prevention program, which we discuss further in a subsequent section. Overall, their results were consistent with theories of stress proliferation in which major life stressors lead to secondary stressors (e.g., increased negative affect and problems in close relationships) that have enduring consequences (Pearlin, Schieman, Fazio, & Meersman, 2005). Although not directly assessed, the authors speculated that processes of peer differential affiliation and perceptions of risk behavior and substance use among peers operate to “carry” the observed effects.

A potential advantage of the methodology employed by Brody et al. (2010) over traditional cross-lagged panel designs is the ability to draw inferences on intraindividual (within-person) change in the construct of interest. That is, Brody et al. were able to demonstrate that paths from life stress, through increased negative emotions, predicted actual increases over time in risk behavior, not just the rank-ordering of youth in the sample on this variable, although their statistical analysis was conducted in two steps (SEM then linked to a separate latent growth curve model). Other researchers interested in developmental cascades have employed similar analytic methods to tap into within-person or within-group change over time in constructs of interest, such as broad externalizing and internalizing problems, interparental conflict, and substance use. For example, in related fashion to the analysis conducted by Brody et al. (2010), Lynne-Landsman, Bradshaw, and Ialongo (2010) used growth mixture modeling to identify separate subgroups of cigarette, alcohol, and marijuana use across high school in a predominantly African American and urban sample. The authors embedded this growth mixture model in a larger mediation model, allowing prediction both from middle school risk factors to high school substance use classes and from high school substance use to young adulthood outcomes. Results suggested that male gender, lower parental education, and earlier oppositional behavior and peer rejection were associated with increasing trajectories of cigarette and marijuana use. Substance-using peer affiliation in middle school mediated effects of earlier predictors, including academic performance, on trajectory of alcohol use in high school. Overall, results are consistent with earlier cascade theories of the development of problem behavior (e.g., Patterson, DeBaryshe, & Ramsey, 1989) and suggest potential differences of effects across class of substance (cigarettes versus alcohol versus marijuana). The latent growth curve model can be applied to more than one construct of interest at a time, in an analysis termed parallel process modeling (Cheong, MacKinnon, & Khoo, 2003). This approach was employed by Kouros, Cummings, and Davies (2010) to test relations among trajectories of interparental conflict and child externalizing problems in a community sample throughout early childhood (ages 6–8). Parallel process growth modeling allows testing of associations between both initial level (intercept) and within-person change over time (slope) across multiple constructs, and Kouros and colleagues showed that slope of interparental conflict predicted slope of externalizing, while the reverse was not the case. In addition, slope of externalizing predicted social competence/prosocial behavior across multiple informants later in childhood (age 12), suggesting that externalizing problems mediated an association between changes in interparental conflict and later social competence. Gender differences were observed in mean levels of constructs of interest but not in associations. This work extends earlier in development the processes between marital conflict and behavior problems previously identified in adolescence by Cui, Conger, and Lorenz (2005), and is notable for operationalizing an environmental risk (interparental conflict) as a dynamic process unfolding over time. Klimes-Dougan and colleagues (2010) provided an intriguing example of how one might employ a growth curve modeling approach to examine cascade-related processes, analyzing relations among externalizing, internalizing, and thought problems in a sample of offspring of healthy versus unipolar depressed versus bipolar depressed mothers (youth assessed at four

time points from approximately ages 5 to 20). Drawing from concepts in systems biology, Klimes-Dougan and colleagues argued that a cascade across domains can be conceptualized as a cumulative process in one domain affecting change in another domain of interest. More specifically, this approach used a cumulative externalizing problems score as a focal predictor of the growth curve of internalizing problems and thought problems (considered separately), with interactions with parental mood disorder status also tested. Key results of this analysis showed evidence for cumulative externalizing problems (rather than concurrent problems) predicting estimated age 20 thought problems and internalizing problems, accounting for change over time in the latter domains. In addition, there was some evidence that the cascade to thought problems was more characteristic of offspring of bipolar mothers while the cascade to internalizing problems was more characteristic of offspring of unipolar depressed mothers. Cascade Research Employing Randomized Controlled Trials Reports from prevention trials often test longitudinal links among constructs of interest within the framework of a randomized controlled trial (RCT), typically by including group assignment as a key exogenous predictor in the model. This analysis is potentially quite powerful, since it blends the causal-inference strengths of an RCT with the statistical controls of cross-lagged panel SEM. For example, McClain and colleagues (2010) tested longitudinal relations among several adolescent outcomes 6 years after families were randomized to an intervention program for middle childhood youth who experienced a recent parental divorce and were living primarily with their mothers (see also Zhou, Sandler, Millsap, Wolchik, & DawsonMcClure, 2008). Results suggested that program participation caused changes in two key parenting constructs—the mother-child relationship and effective discipline—which were then associated longitudinally with different classes of problems. More specifically, an improved mother-child relationship mediated program effects on child internalizing problems, while improved discipline practices mediated program effects on child externalizing problems. Although effects were generally small in magnitude, the impressive 6-year prediction in this study shows promise for theory-driven interventions for youth adjusting to divorce (McClain et al., 2010). Striking evidence for cascading positive effects following successful intervention for childhood behavior problems are discussed in detail by Patterson et al. (2010), who provided an overview of four decades of work from the Oregon Social Learning Center on the development of serious antisocial behavior and aggression in youth: the Parent Management Training—Oregon Model framework. Two points from this work are of particular relevance. First, work from the Oregon group broadly parallels the work by Dodge and colleagues (2009) just described, detailing a progressive sequence in which disrupted early parenting and minorto-moderate forms of child externalizing lead to association with deviant peers, broader ranges of externalizing problems and continued family conflict, eventuating in high risk for adult criminality. Second, Patterson and colleagues described somewhat unexpected and long-lasting positive “spillover” effects into parents' own lives, following successful reduction in child behavior problems through the intervention. More specifically, mothers with children whose problems were reduced reported subsequent improvements in their own social relationships.

This exciting set of findings provides evidence that intervening early in one domain can show systemic effects across other domains. In addition, reported effect sizes from this series of analyses were larger than many seen in nonintervention studies, and in some cases maintained or even increased over time, suggesting that at least for some families, intervention participation created lasting changes in the family context (Patterson et al.). We review additional results from these two research groups later when discussing model prevention programs.

Cascade Models: Summary The work just reviewed shows interesting convergence of broad findings across a variety of domains, samples, and methods. Generally speaking, these empirical investigations confirm the crucial significance of early behavioral regulation and rule-abiding conduct (i.e., lack of externalizing problems/aggression) for transitioning to school and success in the childhood years. At the same time, when such problems affect academic functioning and social relationships, effects can feed back to amplify externalizing difficulties as well as lead to difficulties in the broad internalizing domain (i.e., anxiety and depressive symptoms). Extensions of cascade modeling have noted related empirical findings in domains of EC and self-regulation, alcohol and substance use, and several other areas of interest for developmental psychopathology. Importantly, cascade-related research can also be informative in cases where cross-domain longitudinal links are absent. For example, Tully, Iacono, and McGue (2010) reported analyses from the Minnesota Twin Family Study examining the covariation across ages 15 to 21 of two prominent mental/behavioral disorder symptoms: nicotine dependence and major depressive disorder. Results from this genetically informed study, which also noted increasing heritability of both constructs with age, supported models in which the two sets of symptoms were related only through concurrent covariation, ruling out one domain having a direct effect on the other. The work just reviewed opens methodological and statistical avenues for future research on the interplay of competence and psychopathology over time. In particular, the use of latent growth curve modeling and the embedding of growth models within larger mediation models (Brody et al., 2010; Klimes-Dougan et al., 2010) may allow for more detailed testing of some of the theoretical predictions made in this area of research. More generally, the papers reviewed show that both cascade theories and developmental psychopathology orientations have motivated researchers to expand their statistical and analytic toolboxes, often resulting in intriguing modeling combinations. Although the contribution of these different modeling approaches is expected to vary across future work, it is heartening to see the simultaneous exploration of new theory and new methods. It is also consistent with broader recent calls in psychology for modeling-based inference (Rodgers, 2010), even if models are developed within narrow domains, as well as calls for building analytic models that rule out observations as well as rule them in (Roberts & Pashler, 2000). However, at the current stage of cascade research, it seems clear that additional theoretical work is also crucially needed. That is, the field is in need of stronger theories that make more

precise predictions about how varied domains of competence and psychopathology will interact and affect each other. It is likely that as with so many other domains of social science research, progress in theory development is slowed by a continuing reliance on null hypothesis significance testing, which does not motivate more stringent theoretical predictions (Meehl, 1967; Roisman, Booth-Laforce, Belsky, Burt, & Groh, 2013). One possible way forward is a move toward theories that make rank-order predictions regarding the values of model parameters rather than ones that posit the presence or absence of nonzero parameters. This work will be difficult, since some of the strongest theoretical advances to date have focused on specific domains of functioning across relatively narrow time ranges. However, the integrated work by Patterson, Dodge, and others regarding the chain of effects in the development of violence and other antisocial behavior provides a model for related work in other areas of functioning, as does the work by DeFoe and colleagues (2013) unpacking the chain of effects surrounding contextual disadvantage, low academic achievement, juvenile delinquency, and depressive symptoms. More generally, an increased focus on effect sizes and their interpretation in cascade research is needed. With very few exceptions, the work reviewed above relies on passive, nonexperimental study of target samples. One of the strongest means the field has to test causal theories in development is by intervening in hypothesized processes and observing whether the externally manipulated change in a target process or mediator then changes the (downstream) target outcome and comparing that to a group which did not receive the experimental manipulation. Well-designed and implemented prevention and intervention studies can represent a powerful window of observation on developmental process. We now turn to a review of selected prevention and intervention work that focuses on the bidirectional associations between competence (success in age-salient developmental tasks) and psychopathology.

Interventions to Promote Competence and Reduce Psychopathology In the second edition of this chapter, we argued that preventive interventions are a particularly useful tool for examining the associations between competence and psychopathology, because randomized controlled trials can serve as true experiments in testing hypothesized causal relationships among variables or as tests of developmental theory (Cicchetti & Hinshaw, 2002). Often researchers think of their interventions as “interactions with development” (Coie et al., 1993; Kellam & Rebok, 1992; Tolan, 2002), such that the intervention is altering one developmental process or trajectory—for example, promoting a specific cognitive, emotional, or behavioral skill—in an effort to achieve a different developmental outcome, such as health and well-being or reduced problem behavior or psychopathology. Theories of how and why competence and psychopathology are connected in individual development have important implications for practice and policy, particularly given the twin societal goals of increasing competence and reducing the psychological, social, and economic suffering associated with psychopathology. Empirically supported conceptual models that link healthy adaptation or competence with behavior problems or psychopathology can serve as conceptual guides for

developing and evaluating interventions. We also noted that prevention science was distinct from health promotion or PYD in some important ways, yet the two disciplines seemed to be converging toward a more unified framework for understanding intervention effects (Catalano, Berglund, Ryan, Lonczak, & Hawkins, 2004). Competence has been a central target of prevention efforts for decades, although this focus has not always been acknowledged, with some criticizing prevention science as maintaining a deficit orientation (Benson, 1997; R. M. Lerner & Benson, 2003). Indeed, the 1994 Institute of Medicine report on prevention science, Reducing Risks for Mental Disorders: Frontiers for Preventive Intervention Research, explicitly excluded mental health promotion efforts. Nonetheless, as we noted in our 2006 chapter, competence promotion efforts have been a fundamental aspect of prevention from the time of President Kennedy's efforts to build community mental health (Kennedy, 1963), to President Carter's Task Panel on Prevention (1978) as part of his Commission on Mental Health, through the American Psychological Association's Board of Professional Affairs Task Force on Promotion, Prevention and Intervention Alternatives in Psychology and its efforts to promote prevention through publication of 14 Ounces of Prevention (Price, Cowen, Lorion, & Ramos-McKay, 1988), one of the first compendia of efficacious prevention programs. Of the ten programs described in that document that focused on working with children and adolescents, eight are designed to enhance competence by building specific social, emotional, intellectual, or behavioral skills that are linked with successful adaptation. Four also emphasize building strengths in other social environmental contexts, such as parenting, as a method of influencing youth development. Since our 2006 chapter, there have been several tangible markers of the convergence in approaches to explicitly link a PYD/health promotion framework with a prevention science approach. One of the best examples of this change in philosophy is the name change for one of the premier venues for accessing information about prevention programs. Since 1996, the Center for the Study and Prevention of Violence at the Institute for Behavioral Sciences at the University of Colorado, Boulder has produced the Blueprints for Violence Prevention, a website and list of efficacious prevention programs to combat violence, problem behavior, and deviance. The outcome criteria for programs to be included on the list were rigorous and narrowly focused on violence as an outcome. Over the years, its website and publications have been some of the primary and most well-respected resources for learning about evidencebased programs for both community members and researchers seeking information about prevention programs that work. Blueprints is now funded by the Annie E. Casey Foundation and has extended its focus from problem behavior exclusively to broader developmental outcomes in the areas of behavior, education, emotional well-being, health, and positive relationships. In doing so, the name also has been changed to Blueprints for Healthy Youth Development. The criteria remain rigorous with less than 5 percent of reviewed programs making the list of model or promising programs. This change represents more than just a shift in language; it appears to be an indicator of the broad conceptual shift that we referenced in the second edition. Equally informative over the intervening years has been the continued accumulation of

empirical evidence that preventive interventions work. A rapidly increasing number of interventions to prevent the occurrence of mental, emotional, and behavioral (MEB) disorders have been implemented and evaluated over the past two decades. Many of these have been tested in RCTs, the gold standard of evaluation, and shown to reduce risk factors, increase protective factors, and decrease the rates of MEB disorders and/or substance abuse for individuals receiving the intervention. It is no longer correct to note that there is no evidence that interventions can change the rates of MEB disorders (National Research Council [NRC] & Institute of Medicine [IOM], 2009). The number of efficacious programs has grown substantially, and this increase, coupled with communities' needs to implement high-quality interventions that work, has generated a variety of mechanisms to disseminate the information about effective programs and best practices for implementing them. Many different websites now contain lists of model or promising programs selected by expert panels after review of the literature or compendia of programs with documentation of their relative efficacy available for community organizations to review. Along with progress in documenting the evidence of interventions have been advancements in the conceptual foundations of these interventions; of note is the extent to which the interventions are guided by an understanding that MEB disorders are developmental. That is, prevention science has reoriented toward a focus on the question “What does the child need 1, 3, and 5 years down the line?” (Beardslee, Chien, & Bell, 2011, p. 248) and emphasizes intervening with children and adolescents because of the high rates of MEB disorders seen by age 24 (Kessler et al., 2005). Similarly influential has been a refocusing of the conceptual models that undergird intervention design, implementation, and evaluation toward a developmental framework that focuses on both MEB disorders and aspects of positive health, referred to in the Institute of Medicine report “Preventing Mental, Emotional and Behavioral Disorders Among Young People” as “developmental competencies” (NRC & IOM, 2009). In the most recent Institute of Medicine (IOM) report on prevention (NRC & IOM, 2009), the committee reconsidered the evidence regarding mental health promotion and determined that although it can be distinguished from prevention of mental disorders in many ways, there is considerable overlap in these approaches conceptually and practically, and mental health promotion is an important component of the intervention spectrum (see also Weisz, Sandler, Durlak, & Anton, 2005). This conceptual shift acknowledges that promoting mental health (a goal of youth development programs) and preventing problems (a goal of prevention science) need not be competing endeavors (Catalano, Berglund, Ryan, Lonczak, & Hawkins, 2004). The primary tenets of these two approaches seem to derive from many common roots; they appear to have arrived at similar conclusions about the essentials for helping young people develop successfully, and both now reflect a developmental approach that includes building competence or wellness as a mechanism for preventing problems (Catalano et al., 2004; E. L. Cowen, 1991, 1995; Masten & Coatsworth, 1998; Weissberg & Greenberg, 1998). This advancement appears to return us to a point in which “prevention at its best represents both an effort to foster competence and prevent problems” (Masten & Coatsworth, p. 216). Although the connections are strong at the broadest level of conceptual organization, specific conceptual models still differ across philosophical approaches and even within approaches across

interventions. We note that across these models, competence is conceptualized at different levels of measurement. For example, some models emphasize specific skills (e.g., problem solving, emotional awareness) while others conceptualize competence as broad constructs akin to our developmental task approach. Clear understanding regarding which specific skills or “developmental competencies” are linked directly or indirectly to psychopathology and should be promoted to prevent MEB disorders is at a relatively early stage. In the effort to integrate promotion and prevention, several different conceptual models have been proposed across psychology, social work, and education (e.g., Collaborative for Academic Social and Emotional Learning, 2005; Guerra & Bradshaw, 2008; Haegerich & Metz, 2009; Kia-Keating, Dowdy, Morgan, & Noam, 2011; Kurtines et al., 2008; Schwartz, Pantin, Coatsworth, & Szapocznik, 2007). We highlight three of these models: core competencies, social emotional learning (SEL), and positive youth development (PYD).

Conceptual Models Core Competencies Model One promising framework that has been proposed as a mechanism linking theoretical and empirical work across promotion and prevention is the core competency framework (Guerra & Bradshaw, 2008). Based on a review of the basic developmental and intervention literatures, a panel of experts in developmental and intervention science selected five core competencies as markers of healthy development. These five competencies were selected because they were closely aligned with assets and strengths proposed in the PYD literature (discussed later), were securely grounded in the human development and developmental psychopathology literature, and captured important aspects of promotive and preventive interventions (Guerra & Bradshaw, 2008). Moreover, these competencies showed empirical linkages to a variety of risk behaviors, problem behaviors, or psychopathology. The five proposed core competencies build on a broad developmental model that incorporates concepts of competence and developmental tasks (Havighurst, 1972; Masten & Coatsworth, 1998). However, core competencies are not isomorphic with competence in age-salient developmental tasks. Specifically, the core competencies in this model are: (1) a positive sense of self; (2) self-control; (3) decision-making skills; (4) a moral system of belief; and (5) prosocial connectedness. Each of these core competencies may actually comprise several more specific skills. Positive sense of self reflects three aspects of the sense of self that emerge early in development and remain salient across important points of increased risk for problem behaviors. First is “selfawareness,” which captures a realistic and accurate assessment of one's psychological and behavioral characteristics. This awareness becomes more acute and sophisticated with age such that during adolescence, youth can provide a more refined assessment of who they are and where they are going. Success in adolescence is partially dependent on youth constructing a positive future perspective, clear goal orientation, and positive and realistic possible selves (Cross & Markus, 1991). A positive view of oneself and one's future is positively associated with good developmental outcomes and negatively related to problems. Self-esteem, high

aspirations, and a sense of mastery is related to less frustration in school, better academic outcomes, and less school dropout (Harter, Bresnick, Bouchey, & Whitesell, 1997); youth with a positive future orientation are also less likely to use alcohol (Henry, Swaim, & Slater, 2005; Robbins & Bryan, 2004). Second, agency, or a sense of volition over one's behavior, is critical for development as youth become more able to direct their own development. Having a strong belief in one's ability to control their actions and environment may be critical for success in many developmental domains (Bandura, 2001) but also with regard to preventing specific problem behaviors, such as substance use, through mechanisms such as developing strong peer refusal skills (Schier, Botvin, Diaz, & Griffin, 1999). Third, self-esteem is often used as a marker of PYD, although it is controversial what high self-esteem means and whether it is associated more with positive or negative outcomes (Baumeister, Campbell, Krueger, & Vohs, 2003). Self-esteem seems to be related to indices of well-being and life satisfaction (Diener, 1984), but relatively little evidence suggests that a global sense of self-esteem is promotive (Fearnow-Kenney, Hansen, & McNeal, 2002): rather, domain-specific evaluations may be more important to one's overall functioning (Marsh & O'Mara, 2008). Self-control, or the ability to regulate and manage one's emotions and actions, is the second core competency. Developing effective self-control skills is a critical developmental task that spans childhood and adolescence (Zelazo & Carlson, 2012). Better regulation in young children is associated with fewer problem behaviors, better peer relationships, and better academic achievement. During adolescence, effective self-control (as reflected in planfulness, problem-solving ability, willingness to delay gratification and restrain impulses) reduces risk for problem outcomes, such as aggression/violence, substance use, and criminality (Gottfredson, 2007; Wills, Walker, Mendoza, & Ainette, 2006). Longitudinally, cascade evidence for EC as a mediator of stress-psychopathology relations is emerging, as already discussed. The third core competency is effective decision-making skills, which also increases with age. By adolescence, youth may be capable of coordinating future goals and perspectives with decisions to be made at the present time and are more capable of understanding the linkages between their decisions and consequences (Reyna & Farley, 2006). SIP models have been implemented in a wide variety of longitudinal and intervention studies and indicate that better decision-making skills reduce the likelihood of drinking and smoking (Epstein, Bang, Zhou, & Botvin, 2007) or being aggressive with peers (Crick & Dodge, 1994) and increase the probability of doing well in school (Zins, Weissberg, Wang, & Wahlberg, 2004) and graduating from high school (Hawkins, Catalano, Kosterman, Abbott, & Hill, 1999). A moral system of belief, the fourth proposed core competency, includes judgments about moral issues such as fairness and involves development of aspects of perspective taking and empathy (Guerra, Nucci, & Huesmann, 1994). A sense of morality has been proposed as central to prosocial action and development of a moral identity (Damon, 2004). Moreover, related aspects, such as sense of respect, trustworthiness, caring, and citizenship, are related to this broad competency and are constructs promoted in character education and PYD programs (Benson, 1997; Berkowitz & Beir, 2007). For example, youth who believe that it is morally wrong to drink alcohol tend not to (Beyers, Toumbourou, Catalano, Arthur, & Hawkins, 2004),

youth who are empathic and more prosocial tend not to be aggressive (Pardini, Lochman & Frick, 2003), and youth who are better able to take another's perspective, respond empathically and resolve interpersonal problems effectively are more likely to do well in school as well (Roeser, Eccles, & Sameroff, 2000; Wentzel & Wigfield, 1997). The fifth competency, prosocial connectedness, is broad and encompasses constructs such as attachment and bonding across families, schools, and peers. A variety of conceptual models propose that youth connections across these domains are primary determinants of adjustment. For example, the Social Development Model proposed by Catalano and colleagues (1996) hypothesized that prosocial bonding to parents, schools, and positive peers will reduce the likelihood of problem behaviors, such as delinquency and substance use. When opportunities, involvement, and rewards are in place, youth are more likely to develop strong bonds with positive role models and to internalize the moral messages and positive values that are being taught. This, in turn, reduces the likelihood of poor developmental outcomes, including aggression, substance use, and poor academic performance. The primary contention of this framework is that mastery of the five core social and emotional competencies naturally or through intervention will yield healthy development and reduce the likelihood of problem behavior. This conceptual model helps researchers organize their intervention approaches, provides specific targets for intervention, and helps guide evaluation of an intervention's effects. This model shows marked similarities to the broad developmental models of competence proposed by our prior writing (e.g., Masten & Coatsworth, 1998; Masten et al., 2006). Social Emotional Learning A second model integrating the prevention and promotion perspectives is the SEL model, which has also received a substantial amount of interest and research attention over the past 15 years. Building from a foundation of developmental and intervention research indicating that inability to master core social-emotional competencies can lead to a variety of problem behaviors and social difficulties (Masten & Coatsworth, 1998; Weissberg & Greenberg, 1998), researchers hypothesized that promoting social and emotional learning in schools would enhance health and well-being and reduce the problems of students. Moreover, leaders in the SEL movement contended that schools' roles in promoting healthy development for children and youth include promoting ways that students can behave responsibly and interact well with others socially and emotionally while also building academic skills. The SEL approach is inherently developmental and focuses on helping youth acquire the competencies that allow them to recognize and manage emotions, set and achieve positive goals, appreciate the perspective of others, establish and maintain positive relationships, make responsible decisions, and handle interpersonal situations constructively (Elias et al., 1997). The five key competencies in the SEL framework were developed in a review of theoretical models of emotional intelligence, social emotional competence promotion, SIP, as well as behavior change and learning theories (e.g., the Health Beliefs Model and the Theory of Reasoned Action). The SEL framework shares considerable overlap with the core competencies model, so is described more briefly here.

The framework proposes that a key set of SEL competencies is related to youth health and well-being, academic achievement, and risky or problem behaviors. In establishing their core competency framework, the leaders of the Collaborative for Academic Social and Emotional Learning (CASEL) drew from theories of SEL, theories of behavior change and empirical findings linking children and youth social adjustment with their health outcomes (Payton et al., 2008). The five key SEL competencies are: (1) self-awareness; knowing what one is feeling, having a realistic assessment of one's abilities and a well-grounded sense of self; (2) selfmanagement, meaning handling emotions well, delaying gratification and persevering in the face of obstacles to goals; (3) social awareness, meaning understanding what others are feeling and being able to take their perspective, and maintaining appreciation for and positive interactions with diverse groups; (4) relationship skills, meaning handling emotions in relationships well, maintaining mutually positive relationships, having the ability to negotiate conflict and resist peer pressure; and (5) responsible decision making, making decisions based on accurate appraisal of the information available and the likely consequences of action and taking responsibility for one's decisions (CASEL, 2005). Beyond articulating this list of competencies, the SEL framework also describes characteristics and features of quality SEL programs that are likely to improve overall effectiveness (Durlak, Weissberg, & Pachan, 2010; Payton et al., 2008). Basic features of the program, such as whether the conceptual rationale was sound and clear, whether the program used effective teaching strategies, whether it was coordinated across an entire school with good connections to the community, and whether it provided adequate teacher training and technical support and used high-quality research methods to evaluate its outcome, are included as recommended features of quality programs that will increase the probability that a program will alter the competencies (Payton et al., 2008). Moreover, CASEL contends that interventions that are Sequenced, Active, Focused, and Explicit (SAFE) will be more effective than those that do not have these characteristics. This acronym suggests that for effective programming, skills should be taught in a sequenced manner in which they build on each other over time, curricula should incorporate active learning techniques that provide sufficient time and focused attention to learning the skills at hand, and finally that there are clear and explicit objectives for what is being taught and learned. Positive Youth Development: The Five Cs The Five Cs model was conceptualized and constructed to provide a concise description of PYD and a simple model for youth practitioners that could effectively and efficiently capture outcomes of natural positive development and specific youth development program effects. It was also intended to be more parsimonious than lengthier lists of PYD attributes, such as the 40 developmental assets proposed by the Search Institute (Scales & Leffert, 1999) or a proposed list of ten tasks of adolescent development (Simpson, 2001; Simpson & Roehlkepartain, 2003), although there are clear connections between these and related constructs from resilience research (Masten, 2014). This model highlights five dimensions representing healthy youth development: Caring, Character, Competence, Confidence, and Connections (R. M. Lerner et al., 2005). Together, these dimensions form a holistic overview

of youth development that includes positive personal development, social-emotional skills, moral values, and cognitive competence (Catalano et al., 2004; Moore & Lippman, 2005). Featured protective factors include positive self-concept, perceived competence, youth-parent relationships, and interpersonal skills. In addition, the framework includes multiple aspects of a person, from internal representations (confidence), values and moral compass (character), empathy and concern for others (caring), efficacious behaviors, skills, and knowledge (competence in personal and interpersonal situations), to external networks with others (connections; R. M. Lerner et al., 2005; Pittman, Irby, Tolman, Yohalem, & Ferber, 2003). Several youth studies have utilized the Five Cs model (e.g., R. M. Lerner et al., 2005), and studies in the United States have demonstrated that parents', youth, and youth practitioners' conceptualizations of thriving and positive development can be summarized using the Five Cs framework (King et al., 2005). At the same time, empirical evidence that the Five Cs can be categorized as a second-order factor of PYD has been demonstrated by R. M. Lerner and colleagues (2005). Although this early work focused strongly on understanding the structure of the Five Cs PYD model and how attainment of the Five Cs promoted thriving, or contribution, described as youth active engagement in citizenship (R. M. Lerner, 2009), the broad PYD model hypothesized negative associations between PYD and risk behaviors (R. M. Lerner et al., 2005). In more recent work, Lerner and colleagues (see Bowers, Geldhof, Johnson, Lerner, & Lerner, 2014; J. V. Lerner et al., 2012; R. M. Lerner et al., 2011) have more clearly integrated these associations into their model. This clearer articulation of the processes by which PYD leads to positive outcomes, such as thriving, and is associated with negative outcomes, such as risky problem behavior (Arbeit et al., 2014), parallels the recent conceptual advancements made in prevention science. Substantial similarities exist across these models, even if the language describing their core aspects is different for each. Other broad and more narrowly defined developmental contextual models also exist (e.g., Kia-Keating et al., 2011; Kurtines et al., 2008) that incorporate many similar features to the three described here. Broadly speaking, they all reflect to some degree an individual's ability to manage emotions, gain a clear and positive sense of self, make responsible decisions, and show respect and cooperation in interacting with others. These dimensions are similar to those broad domains described by Masten and Coatsworth (1998). When considered in their entirety, they reflect the developmental perspective articulated by Waters and Sroufe (1983) in defining a competent person as one who is able to generate and coordinate flexible adaptive responses to interact effectively with the environment.

Overview of Interventions Targeting Improvements in Core Competencies The broad integrative models just discussed are drawn from and have inspired intervention studies to demonstrate the associations between the models' core competencies and outcomes of healthy and unhealthy behaviors. In one of the largest meta-analytic studies of SEL interventions to date, Durlak and colleagues (2011) analyzed results from 213 studies involving 270,034 students from grades kindergarten through high school. This study expanded

the already substantial body of research examining effects of school-based interventions (e.g., Durlak & Wells, 1997; Losel & Beelman, 2003; Tobler et al., 2000; S. J. Wilson & Lipsey, 2007) primarily because it included studies published in the preceding decade and focused on both positive social emotional skills and risk/problem outcomes. The primary aim of this analysis was to examine whether SEL programs yield positive effects across social and emotional skills, attitudes toward self and others, positive social behavior, conduct problems, emotional distress, and academic achievement. Results indicated that mean intervention effects (calculated using Hedge's g, a measure similar to Cohen's d) ranged from a high effect size (ES) of .57 for program influence on social emotional skills to a low of .22 on conduct problems and were statistically significant for all six categories of outcomes. Follow-up data, with a median elapsed time of 52 weeks, showed a slight decrease in size of effects, but all remained statistically significant (ES = .11 to .32). Quality of implementation was an important moderator of these overall effects: Programs that were delivered following the SAFE criteria were significantly more likely to obtain stronger effects. Non-SAFE programs failed to show significant effects in all measured outcomes except three—attitudes, conduct problems, and academic performance—and programs that reported implementation problems showed significant effects on only two outcomes, attitudes and conduct problems. These results highlight that high-quality and well-implemented SEL programs can promote development in key social emotional competencies (i.e., promote PYD) and reduce problem behaviors (e.g., prevent conduct problems). Interestingly, the poorly implemented SEL programs may not have influenced competencies but still seemed to have some effect on youth problem behavior. This meta-analysis helps support findings from other analyses and reviews (Durlak, Weissberg, & Pachan, 2010; Greenberg et al., 2003; Weissberg & Greenberg, 1998) indicating that SEL acts as both a promotive and a preventive intervention. More information, however, is needed to support conceptual models that hypothesize change in skills competencies and change in problem outcomes. Mediation, as described, is a critical step in demonstrating these linkages. Next we describe select programs with evidence that promoting competence mediated the effect of program participation on problem behavior reduction or prevention. The largest study to date that evaluates the PYD framework is the 4-H Study of Positive Youth Development (J. V. Lerner et al., 2012). The study is not a randomized controlled trial of a specific intervention curriculum; rather, the 4-H study is a naturalistic longitudinal study of youth, many of whom are participating in various competence promotion activities through 4-H and other organizations. The study started when youth were in fifth grade in 2002–2003 and has collected data through grade 12 from over 7,000 youth in 42 states (R. M. Lerner et al., 2011). Although not a controlled test of youth development program participation, the study has yielded important information about participation in structured youth activities, the promotion of youth competence and well-being, and the prevention of problem behaviors. The strength of the PYD approach has always been on understanding the positive aspects of youth behavior, but the inherent approach to understanding development across time has consistently examined how positive development was associated with problem behaviors (e.g., Jelicic, Bobek, Phelps, Lerner, & Lerner, 2007; R. M. Lerner et al., 2005) and how involvement in youth development programs could affect positive and negative development and their associations (e.g., Mueller et al., 2011).

Findings confirm hypothesized associations among participation, PYD, and negative behaviors but also suggest some interesting complexities (Arbeit et al., 2014; Lewin-Bizan et al., 2010; Mueller et al., 2011; Phelps et al., 2007). When focused on broad constructs in variablecentered analyses, findings from the 4-H study generally support the hypothesized negative association between a second-order PYD factor and problem behaviors such as depression, substance use, and delinquency (Jelicic, Bobek, et al., 2007; Phelps et al., 2007). And when examining specific outcomes rather than broad composites, PYD shows a significant negative association with tobacco use, with marijuana use and sexual initiation for girls, and with hard drug use for both boys and girls (Schwartz et al., 2010). When accounting for individual difference variations in patterns of PYD and problem behaviors over time, however, and when examining associations of different longitudinal patterns, the findings are more complex. Although most youth cluster into stable patterns of high, moderate, or low positive or negative behaviors, some youth show more marked change in these characteristics across time (Mueller et al., 2011; Phelps et al., 2009). Additionally, although youth with high PYD over time tend to also show a pattern of consistently low problem behaviors and depressive symptoms across time, this is not uniform. Some youth with consistently high PYD show a pattern of increasing risk behaviors (substance use and aggressive/delinquent behaviors) around grades 7 to 9 and then declining in grade 10 (Lewin-Bizan et al., 2010). Similarly, although a pattern of increasing or stable high depressive symptoms was generally related to stable low PYD, some youth also showed relatively high levels of PYD and increasing depression. These patterns are similar for youth who participated in youth development programs. Using latent class analysis, Arbeit and colleagues (2014) examined profiles of problem behaviors (e.g., delinquency, substance use, depression) when youth in the 4-H study were in sixth to twelfth grades and tested whether these were associated with the Five Cs. Distinct profiles of problem behaviors emerged (e.g., “low risk,” “alcohol and aggression”) across time points and demonstrated complex associations with the Five Cs. For example, although youth in both the “alcohol and aggression” and “mental health and other risks” groups were using alcohol, their patterns of use differed, and the “alcohol and aggression” group showed higher scores on confidence and competence, while youth in the “mental health and other risks” group scored higher on character and caring. In general, the findings indicate that youth are involved in many out-of-school activities, with sports being the most common (Zarrett, Peltz, Fay, Carrano, Li, & Lerner, 2007), but it is the combination of sports and participation in more structured activities, such as Boys & Girls Clubs and the YMCA, that has the largest association with broad PYD in sixth to eighth graders. Using a matched comparison design, youth who participated in 4-H activities were compared to youth who participated in other out-of-school activities. Results from eighth grade indicated that 4-H youth reported higher levels of PYD, were more likely to be in the optimal PYD group (i.e., high levels across multiple indicators), had greater contribution, and had higher expectations to go to college (R. M. Lerner et al., 2008). These youth also had lower likelihood of expressing depressive symptoms and of showing other risk behaviors. In grade 11, youth who had participated in 4-H activities and for whom longitudinal data were available reported greater contribution, more overall school engagement and academic competence, and lower rates of engaging in health risk behaviors (e.g., drug use, sexual

intercourse; R. M. Lerner et al., 2011). Agans and colleagues (2014) extended this work to examine how patterns of involvement across a variety of activities (breadth of involvement) are associated with positive and problematic outcomes. Results indicated that youth who participated in a broader variety of activities across time generally showed higher levels of the Five Cs and contribution, but youth with high participation also showed higher rates of depressive symptoms. Across time, a pattern of maintaining high involvement in multiple activities was strongly related to higher scores on PYD and less problem behavior. Patterns that showed a decline in involvement across time were associated with greater chances of substance use, depression, and delinquent behaviors. The findings linking program participation to outcomes are not as strong as might be found in a randomized controlled trial that specifically accounts for baseline differences in important individual characteristics and experiences. Often those baseline characteristics may be related to selection into activities and to outcome measures. For example, when modeling how intentional self-regulation skills (e.g., setting goals, planning how to achieve them, and mounting internal and external resources when confronted with barriers to achieving goals) and participation in youth development programs in grades 8 and 9 relate to PYD outcomes (i.e., the Five Cs and contribution) in grade 10, findings show strong associations of intentional selfdevelopment with PYD outcomes and only modest associations between participation and contribution. Participation in grade 8 had a weak association with intentional self-regulation at grade 9, suggesting that these programs may affect PYD at least in part by building selfregulation skills. However, more detailed analyses of the mediating effects are needed.

Promotion and Prevention by Building Nurturing Environments A commonality of the SEL, PYD, and prevention approaches is the emphasis each places on the contextual characteristics necessary to promote positive behavior and prevent problem behavior. Each identifies the important contribution that nurturing environments, whether they be schools, families, or out-of-school programs, have for reducing adverse social conditions, promoting and reinforcing “prosociality” (D. S. Wilson, O'Brien & Sesma, 2009), limiting problem behaviors, and promoting psychological flexibility (Biglan, Flay, Embry, & Sandler, 2012). This approach is clearly seen in SEL interventions due to their emphasis on providing safe, supportive, and caring learning environments to foster a strong bond to school that will in turn facilitate the youth's acquisition of social-emotional skills (CASEL, 2005; Greenberg, 2003). Reviews of PYD programs emphasize characteristics of programs that work, including positive adult-youth bonds, in a safe environment in which youth can build mastery through challenging experiences and feel valued by their communities (e.g., Eccles & Gootman, 2002). Prevention programs also focus on building protective resources in the environment, such as parenting and family skills (Sandler, Schoenfelder, Wolchik, & MacKinnon, 2011) or community mobilization and organization (Oesterle et al., 2010). Schools and families are the two contexts that have been studied more comprehensively in association with PYD and problem behavior. Not surprisingly, these are also the two contexts that are the foci of many prevention and promotion programs (e.g., Greenberg, 2003; Spoth, Kavanagh, & Dishion, 2002). Various reviews and meta-analyses, with a focus on RCTs, have

illustrated the ability of well-conceptualized and well-implemented interventions to influence parenting practices (Kaminski, Valle, Fiese, & Boyle, 2008; Lochman & van den Steenhooven, 2002; Sandler et al., 2012; Taylor & Biglan, 1998) and school contexts (Durlak & Wells, 1997; Durlak et al., 2011; Foxcroft & Tsertsvadze, 2011; Greenberg, 2004; Greenberg, et al., 2003; Weissberg & Greenberg, 1998). Parenting interventions for parents of infants, children, and adolescents have successfully changed ways in which parents provide guidance, support, discipline, and nurturance. Parent- and family-focused interventions are able to change important processes associated with youth positive and negative behaviors: parents' sense of competence (Sanders, Markie-Dadds, Tully, & Bor, 2000), maternal responsiveness (Brotman et al., 2008), parents' use of effective discipline (Tolan, Gorman-Smith, & Henry 2004; Wolchick et al., 2007), positive communication (Kosterman, Hawkins, Spoth, Haggerty, & Zhu 1997), family problem solving (DeGarmo & Forgatch, 2005), parent youth relationship quality (Coatsworth, Duncan, Greenberg, & Nix, 2010; Dishion & Andrews, 1995), and parental monitoring (Coatsworth et al., 2010; DeGarmo & Forgatch, 2005; Soper, Wolchik, Tien, & Sandler, 2010). School-based interventions alter the context of school by creating a safe, caring learning environment, improving classroom management, and building a community in which youth are interested and invested (CASEL, 2005; Elias et al., 1991; Hawkins et al., 2004; Pianta & Hamre, 2009; Schaps, Battistich, & Solomon, 2004). Moreover, as noted earlier, these school-based interventions effectively promote the development of critical social-emotional skills (Durlak et al., 2011). Finally, there is ample evidence from both school-based and family-based interventions to indicate that that these effects persist over time, with some effects from family-focused interventions sustaining for up to 20 years (Botvin & Griffin, 2004; Durlak et al., 2011; Sandler et al., 2012). The same interventions that foster nurturing environments promote healthy youth development and can also influence problem behaviors. Many parenting and family-focused interventions as well as school-based interventions are effective in reducing youth mental, emotional, and behavioral problems (Durlak & Wells, 1997; Durlak et al., 2011; NRC & IOM 2009; Tobler et al., 2000). Effects are evident on a wide range of outcomes, including externalizing problems (e.g., Greenberg, Kusche, Cook, & Quamma, 1997; Sandler et al., 2010; Tolan, Gorman-Smith, Henry, & Schoeny, 2009), internalizing problems (e.g., Cowan, Cowan, & Hemming, 2005; Trudeau, Spoth, & Acevedo, 2007; Wolchick et al., 2002), delinquency (e.g., Botvin, Griffin, & Nichols, 2006; DeGarmo & Forgatch, 2005; Eckenroade et al., 2010) academic engagement and performance (e.g., Elliott, 1997; Oyserman, Bybee, & Terry, 2006; Spoth, Randall, & Shinn, 2008; Stormshak, Connell, & Dishion, 2009; Tolan et al., 2009), substance use (Botvin & Griffin, 2004; Prado et al., 2007; Spoth, Shin, Guyll, & Acevedo, 2006; Wolchik et al. 2002), and risky sexual behaviors (Pantin et al., 2009; Wolchik et al., 2002).

Testing Mediating Processes The true scientific benefit of these kinds of interventions comes when they are conducted in a rigorously employed and rigorously evaluated RCT with a prospective longitudinal design that can capture changing developmental processes and cascades (MacKinnon, 2008). In such studies, mediation analyses can be used to examine whether the intervention affects proximal

processes, which, in turn, affect other positive or negative developmental outcomes. In these instances, the intervention trials can serve as tests of a theoretical model of development (Cicchetti & Hinshaw, 2002), including a theoretical model of how competence and psychopathology are related. Careful attention to what is manipulated is needed, however, as most RCTs of prevention programs change several core constructs at once (Peterson, 2013). Results from mediation analyses of experimental intervention studies are beginning to accrue and yield important information about how interventions operate and how building specific social emotional skills or constructing a nurturing environment can account for an intervention's effect on problem behavior (Biglan et al., 2012; Sandler et al., 2010). Most mediation analyses have investigated processes that operate to produce short-term positive developmental outcomes (12 or fewer months postintervention) for youth. Much less common are analyses testing the mediating processes that produce long-term (e.g., greater than 5 years) intervention effects (Sandler et al., 2011). Yet results from several exemplary programs illustrate how mediating mechanisms can identify the sequential and cascading developmental chains produced by an intervention. Parent Management Training—Oregon Model As discussed in our section on cascade models, Parent Management Training—Oregon Model (PMTOTM) is a set of evidence-based intervention practices that was developed at the Oregon Social Learning Center by Gerald Patterson and colleagues (2010) that focuses on decreasing and reducing the development of antisocial behavior problems in youth. The practices have been packaged in a variety of interventions designed to influence problem behavior in youth whose families are experiencing a variety of social and contextual challenges. For example, Forgatch has applied these to the context of divorced mothers to develop Parenting Through Change, a 14-session intervention designed to increase skill encouragement, limit-setting, problem solving, monitoring, positive parenting, and decrease coercive parenting in the short term and reduce mental health problems in youth in the long term (DeGarmo & Forgatch, 2005; Forgatch & DeGarmo, 1999; Forgatch et al., 2009). PMTO has been adapted and implemented successfully for diverse communities, including native Norweigians (Ogden & AlmlundHagen, 2008) and immigrant Somali and Pakistani parents living in Norway (Bjørknes & Manger, 2013); and parents in Iceland (Sigmarsdóttir, DeGarmo, Forgatch, & Guðmundsdóttir, 2013), the Netherlands (Bekkema, Wiefferink, & Mikilajczak, 2008), and Mexico City (Baumann et al., 2014). This intervention has produced short- and long-term reductions in youth internalizing and externalizing behaviors and long-term effects on more serious delinquency (DeGarmo et al., 2004; Forgatch & DeGarmo, 1999, 2002). Parenting practices were shown to be significant mediators of youth noncompliance, aggression, internalizing and externalizing, teacher-rated delinquency, and appropriate school behaviors 1 and 3 years postintervention (DeGarmo et al., 2004; Forgatch & DeGarmo, 2005). At 9 years postintervention, effects on both teacher-rated delinquency and number of arrests as documented by police records were mediated by both positive parenting (which included both positive interactions and low coercive interactions) and association with deviant peers (Forgatch et al., 2009). These analyses were strong tests of

developmental processes hypothesized by Dishion and Patterson (2006) to contribute to the onset and escalation of antisocial behavior and enhanced understanding of mechanisms of change in the Parenting Through Change intervention. New Beginnings New Beginnings is a theory-based psychosocial parent training and youth coping enhancement intervention for divorced mothers. The intervention was designed to alter parenting processes and enhance youth coping in 9- to 12-year-old children (Wolchick, Sandler, Weiss, & Winslow, 2007; Wolchick et al., 2000). Results have shown that compared to a home study condition, participation in the intervention significantly improved mothers' parenting postintervention and also reduced child mental health problems but did not directly influence youth coping (Wolchik et al., 2000, 2002). At 6-year follow-up, intervention effects were found on youth internalizing and externalizing, alcohol use, drug use, number of sexual partners, and grade point average (Dawson-McClure, Sandler, Wolchik, & Milsap, 2004; Wolchick et al., 2000, 2002). The New Beginnings research team has systematically tested hypothesized mediating mechanisms for this intervention, and analyses have revealed important information about how the program produced both short- and long-term effects on youth functioning. Analyses showed that program effects on adolescents' grade point average at 6-year follow-up were mediated by intervention-produced improvements on maternal effective discipline (Zhou et al., 2008). For youth who were at elevated risk of problems at baseline, program-induced improvement in mother-child relationship quality mediated the intervention effect on adolescents' mental health problems 6 years later (Zhou et al., 2008). Intervention effects on substance use and risky sexual behavior at 6 years postintervention for those youth who were at higher risk at baseline were mediated by program-induced changes in parental monitoring (Soper et al., 2010). RCTs of the New Beginnings program have been conducted primarily with Caucasian samples (around 90%), but as the program enters into a stage of testing program effectiveness through delivery within the court system, adaptations are being made to ensure it is culturally appropriate for a diverse population (Wolchik et al., 2009). As noted in our cascade discussion, mediation analyses have also shown cascading effects on different dimensions of youth functioning. For example, the program produced immediate improvements in mother-youth relationship quality, which in turn was related to decreases in child internalizing 3 to 4 years later, which led to increases in self-esteem and decreases in internalizing disorders 6 years postintervention (McClain et al., 2010). Another theoretically postulated mediating pathway was supported that showed that increases in effective discipline immediately postintervention led to fewer child externalizing problems at 3 to 4 years postintervention and then to decreased externalizing disorders, decreased substance use, and increased academic achievement (McClain et al., 2010). Moreover, although early tests of the intervention indicated no effect on youth coping (Wolchik et al., 2000), additional analyses indicated that program-induced improvements in mother-child relationship quality led to greater youth coping at 6 months and at 6 years postintervention (Velez, Wolchik, Tien, & Sandler, 2011). Altering the nurturing environment had subsequent effects on youth development of effective coping skills. Similarly, analyses revealed cascading effects such that

intervention effects on high-risk youth's educational expectations and job aspirations at 6 years postintervention were mediated by intervention-induced changes in mother-child relationship quality and externalizing and internalizing behaviors (Sigal, Wolchik, Tien, & Sandler, 2012). These carefully constructed analyses yield important information about how the intervention produced both short- and long-term effects on youth adaptive and maladaptive functioning. Moreover, these analyses begin to test models of how changes in parent variables might mediate long-term intervention effects on youth outcomes by maintaining good-quality parenting, changing youth competencies, or changing the transactions between parent and youth (Sandler et al., 2011). Positive Action Positive Action is an evidence-based PYD and social-emotional and character development program delivered in schools that attempts to promote improvements in prosocial skills, selfcontrol, and academic achievement while also reducing negative behaviors (Flay & Alred, 2010). It is a comprehensive, schoolwide prekindergarten-through-twelfth-grade intervention focusing on helping youth take a positive attitude toward the self, which aims to affect their self-concept, physical and intellectual actions, social-emotional skills for managing their behaviors, and positive interactions with others. Outcome analyses from quasi-experimental and experimental trials of the program have demonstrated its effectiveness in modifying overall school quality and enhancing student academic achievement and school involvement while also influencing violence, substance use, and risky sexual behaviors. Separate trials were conducted in Hawaii and Chicago with multiethnic samples (Beets et al., 2009; Lewis et al., 2012; Li et al., 2011; Snyder, Vuchinich, Acock, Washburn, & Flay, 2012). Main effects were generally comparable across studies, and moderation effects by ethnicity were not reported. Mediation analyses of this PYD intervention have demonstrated that program-induced effects on violence, substance use, and sexual activity were mediated by changes in youth positive academic behaviors (setting goals, motivation, problem solving; Snyder et al., 2010, 2013). Focusing on the proximal social-emotional mediators, analyses from data in three separate trials show that the program reduces the decline in positive behaviors (social-emotional and character development behaviors) that occurs with the transition to adolescence (Washburn et al., 2011). Moreover, mediation analyses indicate that the relatively higher social-emotional skills of youth in intervention schools mediate the effects on substance use. This is one of the first studies to illustrate a mediation pathway in a comprehensive, schoolwide socialemotional and character development program bridging the skills taught in the program and problem behavior outcomes.

Moderation of Prevention Effects via Gene–Environment Interaction In addition to studies of mediation, a compelling future direction for successful prevention programs is identifying moderators of program effects on outcome: that is, determining for whom programs work most effectively. Indeed, the search for mediators and moderators is

often complementary, since moderation implies differential effects which may be “carried” via different paths. Although multiple classes of moderators are of interest to those who design and implement prevention programs, work of particular novelty has focused on moderation of prevention outcomes by genetic variation, thus serving as a class of gene–environment interaction (G×E) research termed gene–intervention (G×I) interaction. An early example of this work has focused on the Strong African-American Families (SAAF) and Strong African-American Families-Teen (SAAF-T) programs, aimed at reducing substance use and other risk behaviors among African American youth living in the rural South; data from this project have also contributed to the cascade literature, as noted earlier (Brody et al., 2010). Selected findings from these studies have demonstrated that youth with the 5-HTTLPR short allele variant who were randomized to the prevention program showed risk behavior comparable to control youth (Brody et al., 2009); that a continuous multilocus genetic “risk” variable (markers from DRD4 and GABA-related loci) interacts with program status in predicting an increased frequency of alcohol use for control participants but not prevention participants (Brody, Chen, & Beach, 2013); and that prevention program participation attenuates the association of negative parenting and diminished telomere length, a biomarker associated with aging (Brody, Yu, Beach, & Philibert, 2015). Major advantages of this randomized design are the ability to rule out gene-environment correlation as a competing explanation for differential effects as well as more accurate effect size estimates for different subgroups of youth (Brody, Beach, et al., 2013). Using data from this project and several others, meta-analytic work confirms moderating effects of genotype on randomized prevention and intervention programs (Bakermans-Kranenburg & van IJzendoorn, 2015; van IJzendoorn & Bakermans-Kranenburg, 2015), emphasizing that such genetic moderation can be conceptualized in the framework of differential susceptibility to environmental effects, for better and for worse (Boyce & Ellis, 2005).

Prevention Programs: Summary Because our chapter focus is broad links between competence and psychopathology, our review of prevention research based on competence promotion is of necessity selective. However, from this research, we can note several exciting developments. First, program terminology is evolving so as to emphasize positive aspects of development, strengths, and assets, and linguistic conflicts between prevention, competence promotion, and positive psychology are fading, as is appropriate. In addition, the empirical evidence base on prevention programs is growing, in particular work that includes large sample sizes followed longitudinally over several years and incorporates relatively intensive assessments. Theoretically, we also observe broad similarities and connections among conceptual frameworks such as the Core Competencies model, SEL, and the Five Cs of PYD, and between these frameworks and our earlier discussion of developmental tasks. Each of these models highlights the importance of skills such as self-awareness, empathy, connections to others, and effective decision making in promoting positive development and, therefore, competence in age-salient developmental tasks. Of particular interest and relevance is the growing body of prevention program evaluations that test theoretically specified models of how proximal

variables mediate the positive effects of program participation as well as recent work on genetic and biological moderation of prevention program effects. For mediation, we see similarities across programs and research teams, with parenting practices (e.g., effective discipline and monitoring), positive parent-child relationships as well as youth goal setting, problem solving, and SIP coming to the fore as key mediators in selected studies. Clearly there is more work to be done in this area: In particular, it will be important to have ongoing communication between intervention and naturalistic researchers in terms of the assessment and measurement of some of the constructs. Many, but not all, of the constructs included by prevention scientists have close referents in naturalistic longitudinal work on competence; it will be important for future work to refine taxonomies of these constructs to avoid both unnecessary overlap as well as excessive fragmentation.

Conclusions and Future Directions One notable change in work on competence and psychopathology seen in recent years is the acknowledgment that these two ostensibly separate areas of inquiry are mutually interdependent and mutually influencing; that this basic idea is now often implied rather than explicitly defended is progress for the field. However, we have advanced to the point where more specific claims about influences among competence and psychopathology should be made, and where, as appropriate, potential limitations of newer methodologies are starting to emerge. First, it is clear that the dominance of the DSM system for the classification of mental disorders creates significant challenges for an integrated science of competence and psychopathology. Although explicitly acknowledged as a system with many stakeholders who often have competing interests, and although created from a governing body with input from developmental psychopathologists (but not guided primarily by them), the DSM nonetheless exerts tremendous direct and indirect influence on the questions posed by researchers in this area. This is particularly evident in the work on functional impairment, where work that seeks to test critical alternatives to DSM-based scales remains limited outside of a few select instruments and where efforts to separate out impairment from diagnostic criteria, although admirable from a scientific viewpoint, are likely to be extremely difficult. Although functional impairment encompasses a very broad range of content—and includes domains less often studied in modern developmental task research, such as physical mobility, self-care, and hobbies and leisure activities—it remains important for our consideration. In particular, the inclusion of impairment measures as part of outcome batteries (and, we hope, intermediate assessments) for intervention RCTs (Becker et al., 2011; Fabiano & Pelham, 2009) is a welcome development that could help clarify the mechanisms of action for some of the field's best-documented treatments for youth mental illness. Second, the organizing idea of developmental cascades has motivated a substantial body of recent research on longitudinal links between salient constructs. As noted, the focus of this work has broadened from early studies focusing on major competence domains and broad psychopathology dimensions of internalizing and externalizing, to studies that include a wide

range of other individual and contextual variables, such as effortful control, stressful life events, and divorce. Both early and more recent work have demonstrated the key significance of academics and social competence as broad-based developmental tasks that, in multiple studies, have mediated the longitudinal associations between externalizing and internalizing problems in paths that support, among other theories, dual failure models of the development of depressive symptoms (Capaldi, 1992). Work focused on academics underscores the early development of important linguistic and pre-academic skills and the generally high rank-order continuity of academics throughout childhood and adolescence; nonetheless, multiple studies show evidence for externalizing problems potentially undermining academic achievement and for academic failures potentially contributing to increases in internalizing problems. Academic skills themselves can be predicted all the way from infancy, albeit modestly, through pathways involving physical mobility and exploration (Bornstein et al., 2013c). Work focused on social competence has shown compelling evidence for its mediating role between externalizing and internalizing problems. In addition, multiple reports show evidence for increasingly entangled associations between social functioning and externalizing through middle to late childhood and early adolescence. Although the presence of a close friend can serve to mitigate some of the negative associations between social difficulties and internalizing problems, the nature of friendships must be considered carefully, since deviant peer associations (via deviancy training) are a key driver of escalation from moderate externalizing problems to more serious violent behavior in adolescence (Van Ryzin & Dishion, 2013). Social competence is a particularly difficult target construct and is perhaps better conceptualized as a family of constructs. Associations between social functioning and other competence and problem domains are expected to vary depending on the breadth of assessment and, more specifically, whether peer status and friendship measures are included; they may also vary by context and rater (Dirks, Treat, & Weersing, 2007). For example, a full cascade from externalizing through self-reported peer victimization to internalizing problems was found in early childhood by van Lier and colleagues (2012), but not using a maternal-report measure of broad social competence by Burt and Roisman (2010), suggesting possible differences by either informant or construct breadth. However, we still have relatively little basis on which to theorize differential prediction depending on these varied assessments; thus, further work delineating key aspects of social competence is needed. One intriguing example, related to the NIMH RDoC initiative described earlier, is research by Iarocci, Yager, and Elfers (2007) on G×E interactions in social competence that argues for developing social endophenotypes for molecular genetic study possibly based on well-defined constructs such as sociability, face recognition, emotion recognition, and theory of mind, and the study of extreme groups formed via crossing of these constructs. Very few of the cascade studies reviewed herein assess these more specific facets of social competence. It is important to emphasize that although our general focus on and organization of cascades by academics and social competence represents the primary salience of these two broad domains for childhood adaptation, there is a clear need for cascade-informed work that examines aspects of competence outside of these two areas. Several examples of this broader work have been noted earlier, such as work on EC (Eisenberg et al., 2010) and work on the multifaceted

construct of PYD (Lewin-Bizan et al., 2010). The work we have highlighted from prevention programs provides examples of additional target constructs for future work, such as effective decision making and problem solving in youth, as well as a host of parenting-related variables that have only been rarely studied intensively over time from a cascade framework. There is also a need for careful consideration of neuroscientific, biological, and genetic levels of analysis for future work on competence and psychopathology. Other than the prevention program G×E research discussed and a handful of other studies (e.g., Riggs, Greenberg, Kusché, & Pentz, 2006), the cascade and prevention work reviewed has not included these levels of analysis in structural models. However, blending the multiple-level assessment approach inherent in much biological and molecular genetic studies with the longitudinal and statistically sophisticated work reviewed here is a natural direction for future work. This is especially important given recent interest in the biological embedding of social and environmental experiences—work that tests how both positive and negative environments affect biological processes such as gene expression, promotion of genetic transcription factors, inflammatory immune responses, and other biomarkers of both stress and health (e.g., Boyce, Sokolowski, & Robinson, 2012). Emerging domains of inquiry in population neuroscience (blending demography insights with neuroimaging; Falk et al., 2013) and human social genomics (Slavich & Cole, 2013) highlight new visions of the “social brain” and the potential for cascading effects across levels of analysis that help elucidate precursors and consequences of social relationships as well as new moderators at both behavioral and biological levels. Such work can clarify competence-psychopathology associations in different ways. For example, a recent twin study by Boivin and colleagues (2013) has shown evidence for the short-term longitudinal association between disruptive behaviors in kindergarten and peer problems in first grade being primarily genetically mediated, with controls for continuity in each domain. That is, via mechanisms consistent with evocative gene-environment correlation, genetically influenced disruptive behavior problems appear to lead to negative peer experiences over time. Additionally, as outlined in an earlier section, the NIMH RDoC initiative promises to focus greater research attention on multiple levels of analysis with a particular emphasis on the genetic, molecular, and neurobiological underpinnings of constructs central to this review. The potential further unpacking of social competence is a striking example, but new perspectives in epigenetics and brain-behavior interplay will also influence future causal models of academics, externalizing, internalizing, and many other constructs discussed in this chapter. More generally, we expect that the future will bring more focus on delineating and testing explicit processes at multiple levels of analysis that may account for naturally occurring and intervention-induced connections and cascades among domains of adaptive function encompassed by the broad concepts of competence and psychopathology. Although advances in theory and empirical methods linking competence and psychopathology are clearly evident in the decade of work highlighted in this chapter, much of the theory remains quite general and abstract when it comes to explaining how competence and psychopathology become linked in development. Better specification of change processes will inform intervention as well as theory.

Additional work on measurement of key constructs also is needed. For example, although the questionnaire-based assessments of much internalizing and externalizing research have an extensive history in the field, parsing the boundaries between normal and abnormal functioning in a nuanced way and capturing the complexities of contextual interactions will likely require integration of observational and dyadic methods in addition to well-validated questionnaire and checklist measures. For example, the careful theoretically informed creation of structured observational methods, joined with novel questionnaire measures, has led to potential advances in the refinement of developmental models of disruptive behavior problems in preschoolers (Wakschlag, Choi, et al., 2012; Wakschlag, Henry, et al., 2012). Other general considerations are important for future work on competence and psychopathology. First, as in all other areas of science, replication across independent samples is essential for continued progress (e.g., Dodge et al., 2008, 2009; Patterson et al., 2010). Although several broad findings have now shown replication in the cascades literature, amassing a more detailed body of work will continue to be challenging as research teams across the field have invested tremendous time and resources in longitudinal studies that may or may not contain equivalent assessment measures. Second, we hope that future theoretical and empirical work in this broad domain will include a focus on both effect sizes and potential moderators of observed cross-domain longitudinal associations. We applaud the work by Durlak and colleagues (2011) who have illustrated the value of reporting effect sizes from large-scale meta-analyses in prevention work on SEL and encourage the broader reporting and discussion of effect size. In terms of moderators, it is striking that the large majority of tests of gender moderating cross-domain paths have not found evidence for differences in their strength between males and females. Thus, despite wellknown mean gender differences in many of the constructs studied, the processes that link these domains may be largely similar across gender. However, gender is only one potentially relevant moderator of competence-psychopathology associations, making the careful selection of new moderators a key task of future work. Importantly, increased attention to effect size and a search for moderators are complementary goals, because a moderator is relevant largely to the extent that effect size differences across levels of the moderator exceed a meaningful threshold. Those thresholds will necessarily vary across specific research topics, and further refinement of our theories is needed for their establishment. Third, work on competence and psychopathology will benefit from close consideration of recent work on causal inference and causal mediation (Foster, 2010; Imai et al., 2010; Pearl, 2012). Integration of this technical work will allow for a more explicit consideration of assumptions made in both nonexperimental and experimental work related to causality and will help the field, both by using consistent terminology and also by motivating new research designs and analytic methods that may help us get closer to identifying key causes and intermediary processes that bring about these observed associations. Finally, the field lacks comprehensive evidence on how culture, race, and ethnicity affect relations between competence and psychopathology; this gap is evident across all three major topics of this chapter. In terms of functional impairment, the growing research base on cross-

cultural variation in psychopathology almost uniformly focuses on symptoms and only rarely considers the impairment presumably caused by those symptoms (or contributing to them). Although competence judgments based on developmental tasks can be seen as culturally-bound value judgments, less work in this area has examined processes of acculturation systematically, making this an important future direction (Chen, Fu, & Leng, 2014). In terms of race and ethnicity, several cascade studies reviewed include ethnicity as an analytic covariate, but aside from a few exceptions (e.g., Dishion et al., 2010; Hishinuma et al., 2012; Huang, 2015), these studies do not test for ethnic status as a moderator of key associations. Ideally, future cascade work will test theory-based moderators and include constructs relevant to the development of ethnic minority youth, such as discrimination and ethnic identity, which may elucidate pathways of interest. And although several successful prevention programs have been expanded crossculturally and cross-nationally, the adapting of program content to new countries and cultures raises challenges for testing cultural moderation of program effects. For all three areas of focus, the integration of cultural context and biology through study of gene–culture interplay and cultural epigenetics (Causadias, 2013) will likely be of importance to future research in domains covered in this chapter. It is instructive to consider the concluding remarks from the prior edition of this chapter and assess their focus, impact, and continued relevance (Masten et al., 2006, pp. 725–728). At that time, we celebrated the impact of positive psychology and the PYD movement, along with other work on resilience in youth, that continues to shine light on the competent side of adaptation in general and, in particular, how competence in certain domains at certain times can prevent psychopathology and lead to progressively better outcomes. We also noted the beginnings of complex modeling approaches, such as cascade models, that occupy a more central role in this edition. Although there is much to celebrate in terms of the field's advances in recent years, progress has been (perhaps understandably) slower in integrating mental disorder classification systems such as DSM-5 with broader research on the nature of competence, disorder, impairment, and disability. However, there is great hope for continued progress integrating the study of competence and psychopathology in the near future. We have a growing body of work that is exploring key questions around the definition and assessment of mental disorder and how to best separate out symptoms of disorder from impairment; a set of studies using longitudinal cascade models that collectively help to elucidate pathways by which symptoms and competence interact over time; and a robust and growing body of prevention data that has increasingly employed sophisticated tests of mediation and G×E in addition to well-developed conceptual models. We look forward to increasing integration of all three of these broad areas in the near future, with continued attention to more stringent tests of competing theories in each domain.

References Aalen, O. O., Røysland, K., Gran, J. M., & Ledergerber, B. (2012). Causality, mediation and time: A dynamic viewpoint. Journal of the Royal Statistical Society: Series A (Statistics in Society), 175(4), 831–861. doi: 10.1111/j.1467–985X.2011.01030.x

Agans, J. P., Champine, R. B., DeSouza, L. M., Mueller, M. K., Johnson, S. K., & Lerner, R. M. (2014). Activity involvement as an ecological asset: Profiles of participation and youth outcomes. Journal of Youth and Adolescence, 43, 919–932. doi: 10.1007/s10964–014–0091– 1 American Association on Mental Retardation. (2002). Mental retardation: Definition, classification, and systems of support (10th ed.). Washington, DC: Author. American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.). Arlington, VA: American Psychiatric Publishing. Angold, A., Costello, E. J., & Erkanli, A. (1999a). Comorbidity. Journal of Child Psychology and Psychiatry, 40(1), 57–87. doi: 10.1111/1469–7610.00424 Angold, A., Costello, E. J., Farmer, E. M. Z., Burns, B. J., & Erkanli, A. (1999b). Impaired but undiagnosed. Journal of the American Academy of Child & Adolescent Psychiatry, 38(2), 129–137. doi: 10.1097/00004583–199902000–00011 Ansary, N. S., & Luthar, S. S. (2009). Distress and academic achievement among adolescents of affluence: A study of externalizing and internalizing problem behaviors and school performance. Development and Psychopathology, 21(1), 319–341. doi: 10.1017/S0954579409000182 Arbeit, M. R., Johnson, S. K., Champine, R. B., Greenman, K. N., Lerner, J. V., & Lerner, R. M. (2014). Profiles of problematic behaviors across adolescence: Covariations with indicators of positive youth development. Journal of Youth and Adolescence, 43, 971–990. doi: 10.1007/s10964–014–0092–0 Bailey, J. A., Hill, K. G., Oesterle, S., & Hawkins, J. D. (2006). Linking substance use and problem behavior across three generations. Journal of Abnormal Child Psychology, 34(3), 263–282. doi: 10.1007/s10802–006–9033-z Bailey, J. A., Hill, K. G., Oesterle, S., & Hawkins, J. D. (2009). Parenting practices and problem behavior across three generations: Monitoring, harsh discipline, and drug use in the intergenerational transmission of externalizing behavior. Developmental Psychology, 45(5), 1214–1226. doi: 10.1037/a0016129 Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2015). The hidden efficacy of interventions: Gene × environment experiments from a differential susceptibility perspective. Annual Review of Psychology, 66, 381–409. doi: 10.1146/annurev-psych-010814-015407 Balázs, J., Miklósi, M., Keresztény, Á., Hoven, C. W., Carli, V., Wasserman, C.,… Wasserman, D. (2013). Adolescent subthreshold-depression and anxiety: Psychopathology, functional impairment and increased suicide risk. Journal of Child Psychology and Psychiatry, 54(6), 670–677. doi: 10.1111/jcpp.12016 Bandura, A. (1991). Self-regulation of motivation through anticipatory and self regulatory

mechanisms. In R. A. Dienstbier (Ed.), Perspectives on motivation: Nebraska symposium on motivation (Vol. 38, pp. 79–94). Lincoln, NB: University of Nebraska Press Bandura, A. (1997). Self-efficacy: The exercise of control. New York, NY: Freeman. Barkley, R. A. (2009). Foreword. In S. Goldstein & J. A. Naglieri (Eds.), Assessing impairment: From theory to practice (pp. ix–xi). New York, NY: Springer. Barkley, R. A. (2012a). Barkley Functional Impairment Scale (BFIS). New York, NY: Guilford Press. Barkley, R. A. (2012b). Barkley Functional Impairment Scale—Children and Adolescents (BFIS-CA). New York, NY: Guilford Press. Barkley, R. A. (2012c). Distinguishing sluggish cognitive tempo from attentiondeficit/hyperactivity disorder in adults. Journal of Abnormal Psychology, 121(4), 978–990. doi: 10.1037/a0023961 Barksdale, C. L., Azur, M., & Daniels, A. M. (2010). Behavioral and emotional strengths among youth in systems of care and the effect of race/ethnicity. Journal of Emotional and Behavioral Disorders, 18(4), 236–246. doi: 10.1177/1063426609351700 Baumann, A. A., Rodríguez, M. M. D., Amador, N. G., Forgatch, M. S., & Parra-Cardona, J. R. (2014). Parent Management Training-Oregon model (PMTOTM) in Mexico City: Integrating cultural adaptation activities in an implementation model. Clinical Psychology: Science and Practice, 21(1), 32–47. doi: 10.1111/cpsp.12059 Baumeister, R. F., Campbell, J. D., Krueger, J. I., & Vohs, K. D. (2003). Does high self esteem cause better performance, interpersonal success, happiness, or healthier lifestyles? Psychological Science in the Public Interest, 4, 1–44. doi: 10.1111/1529–1006.01431 Beardslee, W. R., Chien, P. I., & Bell, C. C. (2011). Prevention of mental disorders, substance abuse, and problem behaviors: A developmental perspective. Psychiatric Services, 62, 247– 254. doi: 10.1176/appi.ps.62.3.247 Becker, K. D., Chorpita, B. F., & Daleiden, E. L. (2011). Improvement in symptoms versus functioning: How do our best treatments measure up? Administration and Policy in Mental Health and Mental Health Services Research, 38(6), 440–458. doi: 10.1007/s10488–010– 0332-x Beets, M. W., Flay, B. R., Vuchinich, S., Snyder, F. J., Acock, A., Li, K. K.,… Durlak, J. (2009). Use of a social and character development program to prevent substance use, violent behaviors, and sexual activity among elementary-school students in Hawaii. American Journal of Public Health, 99, 1438–1445. doi: 10.2105/ajph.2008.142919 Bekkema, N., Wiefferink, C., & Mikolajczak, J. (2008). Implementing the Parent Management Training Oregon model in the Netherlands. Emotional and Behavioural Difficulties, 13(4),

249–258. doi: 10.1080/13632750802442136 Benson, P. L. (1997). All kids are our kids: What communities must do to raise caring and responsible children and adolescents. San Francisco, CA: Jossey-Bass. Berkowitz, M. W., & Bier, M. C. (2007). What works in character education? Journal of Research in Character Education, 5, 29–48. Beyers, J. M., Toumbourou, J. W., Catalano, R. F., Arthur, M. W., & Hawkins, J. D. (2004). A cross-national comparison of risk and protective factors for adolescent substance use: The United States and Australia. Journal of Adolescent Health, 35, 3–16. doi: 10.1016/s1054– 139x(03)00344–6 Biglan, A., Flay, B. R., Embry, D. D., & Sandler, I. N. (2012). The critical role of nurturing environments for promoting human well-being. American Psychologist, 67(4), 257–271. doi: 10.1037/a0026796 Binet, A., & Simon, T. (1905). The development of intelligence in children. (E. S. Kite, Trans.). Baltimore, MD: Williams & Wilkins. Bird, H. R., Canino, G. J., Davies, M., Ramírez, R., Chávez, L., Duarte, C., & Shen, S. (2005). The Brief Impairment Scale (BIS): A multidimensional scale of functional impairment for children and adolescents. Journal of the American Academy of Child & Adolescent Psychiatry, 44(7), 699–707. doi: 10.1097/01.chi.0000163281.41383.94 Bjørknes, R., & Manger, T. (2013). Can parent training alter parent practice and reduce conduct problems in ethnic minority children? A randomized controlled trial. Prevention Science, 14(1), 52–63. doi: 10.1007/s11121–012–0299–9 Blandon, A. Y., Calkins, S. D., Grimm, K. J., Keane, S. P., & O'Brien, M. (2010). Testing a developmental cascade model of emotional and social competence and early peer acceptance. Development and Psychopathology, 22(4), 737–748. doi: 10.1017/s0954579410000428 Block, J. H., & Block, J. (1980). The role of ego-control and ego-resiliency in the organization of behavior. In W. A. Collins (Ed.), Development of cognition, affect, and social relations (pp. 39–101). Hillsdale, NJ: Erlbaum. Boivin, M., Brendgen, M., Vitaro, F., Forget-Dubois, N., Feng, B., Tremblay, R. E., & Dionne, G. (2013). Evidence of gene–environment correlation for peer difficulties: Disruptive behaviors predict early peer relation difficulties in school through genetic effects. Development and Psychopathology, 25(1), 79–92. doi: 10.1017/S0954579412000910 Bölte, S., de Schipper, E., Holtmann, M., Karande, S., de Vries, P. J., Selb, M., & Tannock, R. (2014a). Development of ICF Core Sets to standardize assessment of functioning and impairment in ADHD: the path ahead. European Child and Adolescent Psychiatry, 23, 1139– 1148. doi: 10.1007/s00787-013-0496-5

Bölte, S., de Schipper, E., Robison, J. E., Wong, V. C. N., Selb, M., Singhai, N.,… Zwaigenbaum, L. (2014b). Classification of functioning and impairment: The development of ICF Core Sets for autism spectrum disorder. Autism Research, 7(1), 167–172. doi: 10.1002/aur.1335 Bornstein, M. H., Hahn, C.-S., & Haynes, O. M. (2010). Social competence, externalizing, and internalizing behavioral adjustment from early childhood through early adolescence: Developmental cascades. Development and Psychopathology, 22(4), 717–735. doi: 10.1017/S0954579410000416 Bornstein, M. H., Hahn, C.-S., & Suwalsky, J. T. D. (2013a). Developmental pathways among adaptive functioning and externalizing and internalizing behavioral problems: Cascades form childhood into adolescence. Applied Developmental Science, 17(2), 76–87. doi: 10.1080/10888691.2013.774875 Bornstein, M. H., Hahn, C.-S., & Suwalsky, J. T. D. (2013b). Language and internalizing and externalizing behavioral adjustment: Developmental pathways from childhood to adolescence. Development and Psychopathology, 25, 857–878. doi: 10.1017/S0954579413000217 Bornstein, M. H., Hahn, C.-S., & Suwalsky, J. T. D. (2013c). Physically developed and exploratory young infants contribute to their own long-term academic achievement. Psychological Science, 24(10), 1906–1917. doi: 10.1177/0956797613479974 Bornstein, M. H., Hahn, C.-S., & Wolke, D. (2013). Systems and cascades in cognitive development and academic achievement. Child Development, 84(1), 154–162. doi: 10.1111/j.1467–8624.2012.01849.x Bowers, E. P., Geldhof, G. J., Johnson, S. K., Lerner, J. V., & Lerner, R. M. (2014). Special issue introduction: Thriving across the adolescent years. Journal of Youth and Adolescence, 43, 859–868. doi: 10.1007/s10964–014–0117–8 Boyce, W. T., & Ellis, B. J. (2005). Biological sensitivity to context I: An evolutionarydevelopmental theory of the origins and functions of stress reactivity. Development and Psychopathology, 17, 271–301. doi: 10.1017/s0954579405050145 Boyce, W. T., Sokolowski, M. B., & Robinson, G. E. (2012). Toward a new biology of social adversity. PNAS, 109(Suppl. 2), 17143–17148. doi: 10.1073/pnas.1121264109 Breslau, N., & Alvarado, G. F. (2007). The clinical significance criterion in DSM-IV posttraumatic stress disorder. Psychological Medicine, 37(10), 1437–1444. doi: 10.1017/S0033291707000426 Brody, G. H., Beach, S. R. H., Hill, K. G., Howe, G. W., Prado, G., & Fullerton, S. M. (2013). Using genetically informed, randomized prevention trials to test etiological hypotheses about child and adolescent drug use and psychopathology. American Journal of Public Health, 103(Suppl. 1), S19–S24. doi: 10.2105/ajph.2012.301080

Brody, G. H., Beach, S. R. H., Philibert, R. A., Chen, Y.-F., & Murray, V. M. (2009). Prevention effects moderate the association of 5-HTTLPR and youth risk behavior initiation: Gene × environment hypotheses tested via a randomized prevention design. Child Development, 80(3), 645–661. doi: 10.1111/j.1467–8624.2009.01288.x Brody, G. H., Chen, Y.-F., & Beach, S. R. H. (2013). Differential susceptibility to prevention: GABAergic, dopaminergic, and multilocus effects. Journal of Child Psychology and Psychiatry, 54(8), 863–871. doi: 10.1111/jcpp.12042 Brody, G. H., Chen, Y.-F., & Kogan, S. M. (2010). A cascade model connecting life stress to risk behavior among rural African American emerging adults. Development and Psychopathology, 22(3), 667–678. doi: 10.1017/S0954579410000350 Brody, G. H., Yu, T., Beach, S. R. H., & Philibert, R. A. (2015). Prevention effects ameliorate the prospective association between nonsupportive parenting and diminished telomere length. Prevention Science, 16, 171–180. doi: 10.1007/s11121–014–0474–2 Brotman, L. M., Gouley, K. K., Huang, K-.Y., Rosenfelt, A., O'Neal, C., & Shrout, P. (2008). Preventive intervention for preschoolers at high risk for antisocial behavior: Long-term effects on child physical aggression and parenting practices. Journal of Clinical Child and Adolescent Psychology, 37, 386–96. doi: 10.1080/15374410801955813 Bruininks, R., Woodcock, R., Weatherman, R., & Hill, B. (1996). Scales of Independent Behavior–Revised. Chicago, IL: Riverside. Bukowski, W. M., Laursen, B., & Hoza, B. (2010). The snowball effect: Friendship moderates escalations in depressed affect among avoidant and excluded children. Development and Psychopathology, 22(4), 749–757. doi: 10.1017/S095457941000043X Burke, J. D., Loeber, R., Lahey, B. B., & Rathouz, P. J. (2005). Developmental transitions among affective and behavioral disorders in adolescent boys. Journal of Child Psychology and Psychiatry, 46(11), 1200–1210. doi: 10.1111/j.1469–7610.2005.00422.x Burke, J. D., Rowe, R., & Boylan, K. (2014). Functional outcomes of child and adolescent oppositional defiant disorder symptoms in young adult men. Journal of Child Psychology and Psychiatry, 55(3), 264–272. doi: 10.1111/jcpp.12150 Burt, K. B., Obradović, J., Long, J. D., & Masten, A. S. (2008). The interplay of social competence and psychopathology over 20 years: Testing transactional and cascade models. Child Development, 79(2), 359–374. doi: 10.1111/j.1467–8624.2007.01130.x Burt, K. B., & Roisman, G. I. (2010). Competence and psychopathology: Cascade effects in the NICHD Study of Early Child Care and Youth Development. Development and Psychopathology, 22(3), 557–567. doi: 10.1017/S0954579410000271 Capaldi, D. M. (1992). Co-occurrence of conduct problems and depressive symptoms in early adolescent boys: II. A 2-year follow-up at Grade 8. Development and Psychopathology, 4(1),

125–144. doi: 10.1017/s0954579400005605 Card, N. A., & Barnett, M. A. (2015). Methodological considerations in studying individual and family resilience. Family Relations, 64, 120–133. doi: 10.1111/fare.12102 Castro, S., Ferreira, T., Dababnah, S., & Pinto, A. I. (2013). Linking autism measures with the ICF-CY: Functionality beyond the borders of diagnosis and interrater agreement issues. Developmental Neurorehabilitation, 16(5), 321–331. doi: 10.3109/17518423.2012.733438 Catalano, R. F., Berglund, M. L., Ryan, J. A. M., Lonczak, H. S., & Hawkins, J. D. (2004). Positive youth development in the United States: Research findings on evaluations of positive youth development programs. Annals of the American Academy of Political and Social Science, 591, 98–124. doi: 10.1177/0002716203260102 Causadias, J. M. (2013). A roadmap for the integration of culture into developmental psychopathology. Development and Psychopathology, 25, 1375–1398. doi: 10.1017/S0954579413000679 Charlesworth, W. R. (1979). Ethology: Understanding the other half of intelligence. In M. V. Cranach, K. Foppa, W. Lepenies, & Ploog (Eds.), Human ethology: Claims and limits of a new discipline (pp. 491–519). Cambridge, UK: Cambridge University Press. Chen, X., Fu, R., & Leng, L. (2014). Culture and developmental psychopathology. In M. Lewis & K. D. Rudolph (Eds.), Handbook of developmental psychopathology (3rd ed., pp. 225– 241). New York, NY: Springer. doi: 10.1007/978–1–4614–9608–3_12 Chen, X., Huang, X., Chang, L., Wang, L., & Li, D. (2010). Aggression, social competence, and academic achievement in Chinese children: A 5-year longitudinal study. Development and Psychopathology, 22(3), 583–592. doi: 10.1017/S0954579410000295 Chen, X., Huang, X., Wang, L., & Chang, L. (2012). Aggression, peer relationships, and depression in Chinese children: A multiwave longitudinal study. Journal of Child Psychology and Psychiatry, 53(12), 1233–1241. doi: 10.1111/j.1469–7610.2012.02576.x Cheong, J., MacKinnon, D. P., & Khoo, S. T. (2003). Investigation of mediational processes using parallel process latent growth curve modeling. Structural Equation Modeling, 10(2), 238–262. doi: 10.1207/S15328007SEM1002_5 Cicchetti, D. (1984). The emergence of developmental psychopathology. Child Development, 55(1), 1–7. doi: 10.2307/1129830 Cicchetti, D. (2006). Development and psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology (2nd ed., Vol. 1, pp. 1–23). New York, NY: Wiley. Cicchetti, D. (2013). An overview of developmental psychopathology. In P. D. Zelazo (Ed.), The Oxford handbook of developmental psychology (Vol. 2, pp. 455–480). New York, NY: Oxford University Press.

Cicchetti, D., & Hinshaw, S. P. (2002). Editorial: Prevention and intervention science: Contributions to developmental theory. Development and Psychopathology, 14, 667–671. doi: 10.1017/s0954579402004017 Cicchetti, D., & Rogosch, F. A. (1996). Equifinality and multifinality in developmental psychopathology. Development and Psychopathology, 8(4), 597. doi: 10.1017/s0954579400007318 Cicchetti, D., & Schneider-Rosen, K. (1986). An organizational approach to childhood depression. In M. Rutter, C. Izard, & P. Read (Eds.), Depression in young people: Clinical and developmental perspectives (pp. 71–134). New York, NY: Guilford Press. Cleverley, K., Bennett, K., & Duku, E. (2013). Effects of functional impairment on internalizing symptom trajectories in adolescence: A longitudinal, growth curve modelling study. Journal of Adolescence, 36(1), 45–53. doi: 10.1016/j.adolescence.2012.09.001 Coatsworth, J. D., Duncan, L. G., Greenberg, M. T., & Nix, R. L. (2010). Changing parents' mindfulness, child management skills, and relationship quality with their youth: Results from a randomized pilot intervention trial. Journal of Child and Family Studies, 19, 203–217. doi: 10.1007/s10826–009–9304–8 Coffman, D. L. (2011). Estimating causal effects in mediation analysis using propensity scores. Structural Equation Modeling, 18(3), 357–369. doi: 10.1080/10705511.2011.582001 Coghill, D., & Sonuga-Barke, E. J. S. (2012). Annual research review: Categories versus dimensions in the classification and conceptualisation of child and adolescent mental disorders —implications of recent empirical study. Journal of Child Psychology and Psychiatry, 53(5), 469–489. doi: 10.1111/j.1469–7610.2011.02511.x Coie, J. D., Watt, N. F., West, S. G., Hawkins, J. D., Asarnow, J. R., Markman, H. J.,… Long, B. (1993). The science of prevention: A conceptual framework and some directions for a national research program. American Psychologist, 48, 1013–1022. doi: 10.1037/0003– 066x.48.10.1013 Cole, D. A., Martin, J. M., Powers, B., & Truglio, R. (1997). A competency-based model of child depression: A longitudinal study of peer, parent, teacher, and self-evaluations. Journal of Child Psychology and Psychiatry, 38, 505–514. doi: 10.1111/j.1469–7610.1997.tb01537.x Cole, D. A., & Maxwell, S. E. (2003). Testing mediational models with longitudinal data: Questions and tips in the use of structural equation modeling. Journal of Abnormal Psychology, 112(4), 558–577. doi: 10.1037/0021–843X.112.4.558 Cole, D. A., & Maxwell, S. E. (2009). Statistical methods for risk-outcome research: Being sensitive to longitudinal structure. Annual Review of Clinical Psychology, 5(1), 71–96. doi: 10.1146/annurev-clinpsy-060508–130357 Collaborative for Academic, Social, and Emotional Learning. (2005). Safe and sound: An

educational leader's guide to evidence-based social and emotional learning programs— Illinois edition. Chicago, IL: Author. Cowan, C. P., Cowan, P. A., & Heming, G. (2005). Two variations of a preventive intervention for couples: Effects on parents and children during the transition to elementary school. In P. A. Cowan, C. P. Cowan, J. C. Ablow, V. K. Johnson, & J. R. Measelle (Eds.), The family context of parenting in children's adaptation to elementary school (pp. 227–312). Mahwah, NJ: Erlbaum. Cowen, E. L. (1991). In pursuit of wellness. American Psychologist, 46, 404–408. doi: 10.1037/0003–066x.46.4.404 Cowen, E. L. (1994). The enhancement of psychological wellness: Challenges and opportunities. American Journal of Community Psychology, 24, 239–245. doi: 10.1007/bf02506861 Cox, M. J., Mills-Koonce, R., Propper, C., & Gariépy, J.-L. (2010). Systems theory and cascades in developmental psychopathology. Development and Psychopathology, 22(3), 497– 506. doi: 10.1017/S0954579410000234 Crick, N. R., & Dodge, K. A. (1994). A review and reformulation of social informationprocessing mechanisms in children's social adjustment. Psychological Bulletin, 115, 74–101. doi: 10.1037/0033–2909.115.1.74 Cross, S., & Markus, H. (1991). Possible selves across the life span. Human Development, 34, 230–255. doi: 10.1159/000277058 Crowley, M. D., Coffman, D. L., Feinberg, M. E., Greenberg, M. T., & Spoth, R. L. (2014). Evaluating the impact of implementation factors on family-based prevention programming: methods for strengthening causal inference. Prevention Science, 15(2), 246–255. doi: 10.1007/s11121–012–0352–8 Cui, M., Conger, R. D., & Lorenz, F. O. (2005). Predicting change in adolescent adjustment from change in marital problems. Developmental Psychology, 41(5), 812–823. doi: 10.1037/0012–1649.41.5.812 Cuthbert, B. N., & Kozak, M. J. (2013). Constructing constructs for psychopathology: The NIMH Research Domain Criteria. Journal of Abnormal Psychology, 122(3), 928–937. doi: 10.1037/a0034028 Damon, W. (2004). What is positive youth development? Annals of the American Academy of Political and Social Science, 591, 13–24. doi: 10.1177/0002716203260092 Dawson-McClure, S., Sandler, I. N., Wolchik, S. A., & Millsap, R. E. (2004). Risk as a moderator of the effects of prevention programs for children from divorced families: A sixyear longitudinal study. Journal of Abnormal Child Psychology, 32, 175–190. doi: 10.1023/b:jacp.0000019769.75578.79

Defoe, I. N., Farrington, D. P., & Loeber, R. (2013). Disentangling the relationship between delinquency and hyperactivity, low achievement, depression, and low socioeconomic status: Analysis of repeated longitudinal data. Journal of Criminal Justice, 41(2), 100–107. doi: 10.1016/j.jcrimjus.2012.12.002 DeGarmo, D. S., & Forgatch, M. S. (2005). Early development of delinquency within divorced families: Evaluating a randomized preventive intervention trial. Developmental Science, 8, 229–239. doi: 10.1111/j.1467–7687.2005.00412.x DeGarmo, D. S., Patterson, G. R., & Forgatch, M. S. (2004). How do outcomes in a specified parent training intervention maintain or wane over time? Prevention Science, 5, 73–89. doi: 10.1023/b:prev.0000023078.30191.e0 de Schipper, E., Lundequist, A., Wilteus, A. L., Coghill, D., de Vries, P. J., Granlund, M., …, Bölte, S. (2015). A comprehensive scoping review of ability and disability in ADHD using the International Classification of Functioning, Disability and Health-Children and Youth Version (ICF-CY). European Child and Adolescent Psychiatry, 24, 859–872. doi: 10.1007/s00787015-0727-z Diener, E. (1984). Subjective well-being. Psychological Bulletin, 95, 542–575. doi: 10.1037/0033–2909.95.3.542 Dirks, M. A., Treat, T. A., & Weersing, V. R. (2007). Integrating theoretical, measurement, and intervention models of youth social competence. Clinical Psychology Review, 27, 327–347. doi: 10.1016/j.cpr.2006.11.002 Dishion, T. J., & Patterson, G. R. (2006). The development and ecology of antisocial behavior. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology: Vol. 3. Risk, disorder, and adaptation (2nd ed., pp. 503–541). Hoboken, NJ: Wiley. Dishion, T. J., Véronneau, M.-H., & Myers, M. W. (2010). Cascading peer dynamics underlying the progression from problem behavior to violence in early to late adolescence. Development and Psychopathology, 22(3), 603–619. doi: 10.1017/S0954579410000313 Ditterline, J., & Oakland, T. (2009). Relationships between adaptive behavior and impairment. In S. Goldstein & J. A. Naglieri (Eds.), Assessing impairment: From theory to practice (pp. 31–48). New York, NY: Springer Science. Doan, S. N., Fuller-Rowell, T. E., & Evans, G. W. (2012). Cumulative risk and adolescent's internalizing and externalizing problems: The mediating roles of maternal responsiveness and self-regulation. Developmental Psychology, 48(6), 1529–1539. doi: 10.1037/a0027815 Dodge, K. A., Godwin, J., & the Conduct Problems Prevention Research Group. (2013). Social-information-processing patterns mediate the impact of preventive intervention on adolescent antisocial behavior. Psychological Science, 24(4), 456–465. doi: 10.1177/0956797612457394

Dodge, K. A., Greenberg, M. T., Malone, P. S., & Group, C. P. P. R. (2008). Testing an idealized dynamic cascade model of the development of serious violence in adolescence. Child Development, 79(6), 1907–1927. doi: 10.1111/j.1467–8624.2008.01233.x Dodge, K. A., Lansford, J. E., Burks, V. S., Bates, J. E., Pettit, G. S., Fontaine, R., & Price, J. M. (2003). Peer rejection and social information-processing factors in the development of aggressive behavior problems in children. Child Development, 74(2), 374–393. doi: 10.1111/1467–8624.7402004 Dodge, K. A., Malone, P. S., Lansford, J. E., Miller, S., Pettit, G. S., & Bates, J. E. (2009). A dynamic cascade model of the development of substance-use onset. Monographs of the Society for Research in Child Development, 74(3), 1–120. doi: 10.1111/j.1540– 5834.2009.00528 Drabick, D. A. G. (2009). Can a developmental psychopathology perspective facilitate a paradigm shift toward a mixed categorical–dimensional classification system? Clinical Psychology: Science and Practice, 16(1), 41–49. doi: 10.1111/j.1468–2850.2009.01141.x Duncan, G. J., Dowsett, C. J., Claessens, A., Magnuson, K., Huston, A. C., Klebanov, P.,… Japel, C. (2007). School readiness and later achievement. Developmental Psychology, 43(6), 1428–1446. doi: 10.1037/0012–1649.43.6.1428 Durlak, J. A., Weissberg, R. P., & Pachan, M. (2010). A meta-analysis of after-school programs that seek to promote personal and social skills in children and adolescents. American Journal of Community Psychology, 45, 294–309. doi: 10.1007/s10464–010–9300– 6 Durlak, J. A., & Wells, A. M. (1997). Primary prevention mental health programs for children and adolescents: A meta-analytic review. American Journal of Community Psychology, 25, 115–152. doi: 10.1023/A:1024654026646 Eccles, J. S., & Gootman, J. A. (2002). Community programs to promote youth development. Washington, DC: National Academies Press. Eckenrode, J., Campa, M., Luckey, D. W., Henderson, C. R. Jr., Cole, R., Kitzman, H.,… Olds, D. (2010). Long-term effects of prenatal and infancy nurse home visitation on the life course of youths: 19-year follow-up of a randomized trial. Archives of Pediatric and Adolescent Medicine, 164(1), 9–15. doi: 10.1001/archpediatrics.2009.240 Eisenberg, N., Spinrad, T. L., Eggum, N. D., Silva, K. M., Reiser, M., Hofer, C.,… Michalik, N. (2010). Relations among maternal socialization, effortful control, and maladjustment in early childhood. Development and Psychopathology, 22(3), 507–525. doi: 10.1017/S0954579410000246 Elias, M. J., Zins, J. E., Weissberg, R. P., Frey, K. S., Greenberg, M. T., Haynes, N. M.,… Schriver, T. (1997). Promoting social and emotional learning: Guidelines for educators.

Alexandria, VA: Association for Supervision and Curriculum Development. Englund, M. M., & Siebenbruner, J. (2012). Developmental pathways linking externalizing symptoms, internalizing symptoms, and academic competence to adolescent substance use. Journal of Adolescence, 35, 1123–1140. doi: 10.1016/j.adolescence.2012.03.004 Epstein, J. A., Zhou, X. K., Bang, H., & Botvin, G. J. (2007). Do competence skills moderate the impact of social influences to drink and perceived social benefits of drinking on alcohol use among inner-city adolescents? Prevention Science, 8, 65–73. doi: 10.1007/s11121–006– 0054–1 Erikson, E. H. (1968). Identity, youth and crisis. New York, NY: Norton. Erskine, H. E., Ferrari, A. J., Polanczyk, G. V., Moffitt, T. E., Murray, C. J. L., Vos, T.,… Scott, J. G. (2014). The global burden of conduct disorder and attention-deficit/hyperactivity disorder in 2010. Journal of Child Psychology and Psychiatry, 55(4), 328–336. doi: 10.1111/jcpp.12186 Ezpeleta, L., Reich, W., & Granero, R. (2009). Assessment of distress caused by psychopathology in children and adolescents. Escritos de Psicología, 2(2), 19–27. Fabiano, G. A., & Pelham, W. E. (2009). Impairment in children. In S. Goldstein & J. A. Naglieri (Eds.), Assessing impairment: From theory to practice (pp. 105–119). New York, NY: Springer Science. Falk, E. B., Hyde, L. W., Mitchell, C., Faul, J., Gonzalez, R., Heitzeg, M. M.,… Schulenberg, J. (2013). What is a representative brain? Neuroscience meets population science. PNAS, 110(44), 17615–17622. doi: 10.1073/pnas.1310134110 Fanti, K. A., & Henrich, C. C. (2010). Trajectories of pure and co-occurring internalizing and externalizing problems from age 2 to age 12: Findings from the National Institute of Child Health and Human Development Study of Early Child Care. Developmental Psychology, 46(5), 1159–1175. doi: 10.1037/a0020659 Fearnow-Kenney, M., Hansen, W. B., & McNeal, R. B. (2002). Comparison of psychosocial influences on substance use in adolescents: Implications for prevention programming. Journal of Child and Adolescent Substance Abuse, 11, 1–24. doi: 10.1300/j029v11n04_01 First, M. B., & Wakefield, J. C. (2010). Defining “mental disorder” in DSM-V. Psychological Medicine, 40(11), 1779–1782. doi: 10.1017/S0033291709992339 Flay, B. R., & Allred, C. G. (2010). The Positive Action Program: Improving academics, behavior and character by teaching comprehensive skills for successful learning and living. In T. Lovat & R. Toomey (Eds.), International handbook on values education and student wellbeing (pp. 471–501). Dordrecht, the Netherlands: Springer. Foley, D. L., Rowe, R., Maes, H., Silberg, J., Eaves, L., & Pickles, A. (2008). The

relationship between separation anxiety and impairment. Journal of Anxiety Disorders, 22(4), 635–641. doi: 10.1016/j.janxdis.2007.06.002 Forgatch, M. S., & DeGarmo, D. S. (1999). Parenting Through Change: An effective parenting intervention for single mothers. Journal of Consulting and Clinical Psycholology, 67, 711– 724. doi: 10.1037/0022–006x.67.5.711 Forgatch, M. S., & DeGarmo, D. S. (2002). Extending and testing the social interaction learning model with divorce samples. In J. B. Reid, R. G. Patterson, & J. J. Snyder (Eds.), Antisocial behavior in children and adolescents: A developmental analysis and model for intervention (pp. 235–256). Washington, DC: American Psychological Association. Forgatch, M. S., Patterson, G. R., DeGarmo, D. S., & Beldavs, Z. G. (2009). Testing the Oregon delinquency model with 9-year follow-up of the Oregon Divorce Study. Development and Psychopathology, 21, 637–660. doi: 10.1017/s0954579409000340 Foster, E. M. (2010). Causal inference and developmental psychology. Developmental Psychology, 46(6), 1454–1480. doi: 10.1037/a0020204 Francis, S. E., Ebesutani, C., & Chorpita, B. F. (2010). Differences in levels of functional impairment and rates of serious emotional disturbance between youth with internalizing and externalizing disorders when using the CAFAS or GAF to assess functional impairment. Journal of Emotional and Behavioral Disorders, 20(4), 226–240. doi: 10.1177/1063426610387607 Franklin, J. C., Jamieson, J. P., Glenn, C. R., & Nock, M. K. (2015). How developmental psychopathology theory and research can inform the Research Domain Criteria (RDoC) project. Journal of Clinical Child and Adolescent Psychology, 44(2), 280–290. doi: 10.1080/15374416.2013.873981 Gadow, K. D., Kaat, A. J., & Lecavalier, L. (2013). Relation of symptom-induced impairment with other illness parameters in clinic-referred youth. Journal of Child Psychology and Psychiatry, 54(11), 1198–1207. doi: 10.1111/jcpp.12077 Gan, S.-M., Tung, L.-C., Yeh, C.-Y., & Wang, C.-H. (2013). ICF-CY based assessment tool for children with autism. Disability and Rehabilitation, 35(8), 678–685. doi: 10.3109/09638288.2012.705946 Garber, J., & Cole, D. A. (2010). Intergenerational transmission of depression: A launch and grow model of change across adolescence. Development and Psychopathology, 22(4), 819– 830. doi: 10.1017/S0954579410000489 Gathje, R. A., Lewandowski, L. J., & Gordon, M. (2008). The role of impairment in the diagnosis of ADHD. Journal of Attention Disorders, 11(5), 529–537. doi: 10.1177/1087054707314028 Gleason, K., & Coster, W. (2012). An ICF-CY-based content analysis of the Vineland Adaptive

Behavior Scales–II. Journal of Intellectual and Developmental Disability, 37(4), 285–293. doi: 10.3109/13668250.2012.720675 Gonzalez, R. (2009). Transactions and statistical modeling: Developmental theory wagging the statistical tail. In A. Sameroff (Ed.), The transactional model of development: How children and contexts shape each other (pp. 223–246). Washington, DC: American Psychological Association. Gooren, E. M. J. C., van Lier, P. A. C., Stegge, H., Terwogt, M. M., & Koot, H. M. (2011). The development of conduct problems and depressive symptoms in early elementary school children: The role of peer rejection. Journal of Clinical Child & Adolescent Psychology, 40(2), 245–253. doi: 10.1080/15374416.2011.546045 Gottesman, I. I., & Gould, T. D. (2003). The endophenotype concept in psychiatry: Etymology and strategic intentions. American Journal of Psychiatry, 160(4), 636–645. doi: 10.1176/appi.ajp.160.4.636 Gottlieb, G. (2007). Probabilistic epigenesis. Developmental Science, 10(1), 1–11. doi: 10.1111/j.1467–7687.2007.00556.x Greenberg, M. T. (2004). Current and future challenges in school-based prevention: The researcher perspective. Prevention Science, 5, 5–13. doi: 10.1023/b:prev.0000013976.84939.55 Greenberg, M. T., Weissberg, R. P., O'Brien, M. U., Zins, J. E., Fredericks, L., Resnik, H., & Elias, M. J. (2003). Enhancing school-based prevention and youth development through coordinated social, emotional, and academic learning. American Psychologist, 58, 466–474. doi: 10.1037/0003–066x.58.6–7.466 Greenspan, S. I. (1981). Psychopathology and adaptation in infancy and early childhood: Principles of clinical diagnosis and preventive intervention. New York, NY: International Universities Press. Guerra, N. G., & Bradshaw, C. (2008). Core competencies to prevent problem behaviors and promote positive youth development. New Directions in Child and Adolescent Development, 122, 1–17. Guerra, N. G., Nucci, L., & Huesmann, L. R. (1994). Moral cognition and childhood aggression. In L. R. Huesmann (Ed.), Current perspectives in aggressive behavior (pp. 13– 33). New York, NY: Plenum Press. Haegerich,T. M., & Metz, E. (2009). The social and character development research program: Development, goals, and opportunities. Journal of Research in Character Education, 7, 1–20. Haller, M., Handley, E., Chassin, L., & Bountress, K. (2010). Developmental cascades: Linking adolescent substance use, affiliation with substance use promoting peers, and academic achievement to adult substance use disorders. Development and Psychopathology,

22(4), 899–916. doi: 10.1017/S0954579410000532 Hamilton, B. A., Naismith, S. L., Scott, E. M., Purcell, S., & Hickie, I. B. (2011). Disability is already pronounced in young people with early stages of affective disorders: Data from an early intervention service. Journal of Affective Disorders, 131(1–3), 84–91. doi: 10.1016/j.jad.2010.10.052 Harter, S. (1978). Effectance motivation reconsidered: Toward a developmental model. Human Development, 21, 34–64. doi: 10.1159/000271574 Harter, S., Bresnick, S., Bouchey, H., & Whitesell, N. R. (1997). The development of multiple role-related selves in adolescence. Development and Psychopathology, 9, 835–854. doi: 10.1017/s0954579497001466 Havighurst, R. J. (1972). Developmental tasks and education (3rd ed.). New York, NY: David McKay. Hawkins, J. D., Catalano, R. F., Kosterman, R., Abbott, R., & Hill, K. G. (1999). Preventing adolescent health-risk behaviors by strengthening protection during childhood. Archives of Pediatric Adolescent Medicine, 153, 226–234. doi: 10.1001/archpedi.153.3.226 Hawkins, J. D., Smith, B. H., & Catalano, R. F. (2004). Social development and social and emotional learning. In J. E. Zins, R. P. Weissberg, M. C. Wang, & H. J. Walberg (Eds.), Building academic success on social and emotional learning: What does the research say? (pp. 135–150). New York, NY: Teachers College Press. Henry, K. L., Swaim, R. C., & Slater, M. D. (2005). Intraindividual variability of school bonding and adolescents' beliefs about the effect of substance use on future aspirations. Prevention Science, 6, 101–112. doi: 10.1007/s11121–005–3409–0 Herrenkohl, T. I., Kosterman, R., Mason, W. A., Hawkins, J. D., McCarty, C. A., & McCauley, E. (2010). Effects of childhood conduct problems and family adversity on health, health behaviors, and service use in early adulthood: Tests of developmental pathways involving adolescent risk taking and depression. Development and Psychopathology, 22(3), 655–665. doi: 10.1017/S0954579410000349 Hishinuma, E. S., Chang, J. Y., McArdle, J. J., & Hamagami, F. (2012). Potential causal relationship between depressive symptoms and academic achievement in the Hawaiian High Schools Health Survey using contemporary longitudinal latent variable change models. Developmental Psychology, 48(5), 1327. doi: 10.1037/a0026978 Huang, C. (2015). Academic achievement and subsequent depression: A meta-analysis of longitudinal studies. Journal of Child and Family Studies, 24, 434–442. doi: 10.1007/s10826–013–9855–6 Iarocci, G., Yager, J., & Elfers, T. (2007). What gene–environment interactions can tell us about social competence in typical and atypical populations. Brain and Cognition, 65(1),

112–127. doi: 10.1016/j.bandc.2007.01.008 Imai, K., Keele, L., & Tingley, D. (2010). A general approach to causal mediation analysis. Psychological Methods, 15(4), 309–334. doi: 10.1037/a0020761 Insel, T. R., & Cuthbert, B. N. (2009). Endophenotypes: Bridging genomic complexity and disorder heterogeneity. Biological Psychiatry, 66(11), 988–989. doi: 10.1016/j.biopsych.2009.10.008 Insel, T. R., Cuthbert, B., Garvey, M., Heinssen, R., Pine, D. S., Quinn, K.,… Wang, P. (2010). Research domain criteria (RDoC): Toward a new classification framework for research on mental disorders. American Journal of Psychiatry, 167(7), 748–751. doi: 10.1176/appi.ajp.2010.09091379 Jayawickreme, N., Jayawickreme, E., Atanasov, P., Goonasekera, M. A., & Foa, E. B. (2012). Are culturally specific measures of trauma-related anxiety and depression needed? The case of Sri Lanka. Psychological Assessment, 24(4), 791–800. doi: 10.1037/a0027564 Jelicic, H., Bobek, D. L., Phelps, E., Lerner, R., & Lerner, J. V. (2007). Using positive youth development to predict contribution and risk behaviors in early adolescence: Findings from the first two waves of the 4-H Study of Positive Youth Development. International Journal of Behavioral Development, 31, 263–273. doi: 10.1177/0165025407076439 Jelicic, H., Theokas, C., Phelps, E., & Lerner, R. M. (2007). Conceptualizing and measuring the context within person←→context models of human development: Implications for theory, research, and application. In T. D. Little, J. A. Bovaird, & N. A. Card (Eds.), Modeling contextual effects in longitudinal studies (pp. 437–456). Mahwah, NJ: Erlbaum. Kaminski, J. W., Valle, L. A., Filene, J. H., & Boyle, C. L. (2008). A meta-analytic review of components associated with parent training program effectiveness. Journal of Abnormal Child Psychology, 36, 567–89. doi: 10.1007/s10802–007–9201–9 Kellam, S. G., & Rebok, G. W. (1992). Building developmental and etiological theory through epidemiologically based preventive intervention trials. In J. McCord & R. E. Tremblay (Eds.), Preventing antisocial behavior: Interventions from birth through adolescence (pp. 162– 194). New York, NY: Guilford Press. Kennedy, J. F. (1963). Mental illness and mental retardation: Message from the President of the United States relative to mental illness and mental retardation. American Psychologist, 18(6), 280–289. doi: 10.1037/h0044231 Kerr, M., Tremblay, R. E., Pagani, L., & Vitaro, F. (1997). Boys' behavioral inhibition and the risk of later delinquency. Archives of General Psychiatry, 54(9), 809–816. doi: 10.1001/archpsyc.1997.01830210049005 Kessler, R. C., Berglund, P., Demler, O., Jin, K., Merikangas, K. R., & Walters, E. E. (2005). Lifetime prevalence and age-of-onset distributions of DSM-IV disorders in the National

Comorbidity Survey Replication. Archives of General Psychiatry, 62, 593–602. doi: 10.1001/archpsyc.62.6.593 Kia-Keating, M., Dowdy, E., Morgan, M. L., & Noam, G. G. (2011). Protecting and promoting: An integrative conceptual model for healthy development of adolescents. Journal of Adolescent Health, 48, 220–228. doi: 10.1016/j.jadohealth.2010.08.006 King, P. E., Dowling, E. M., Mueller, R. A., White, K., Schultz, W., Osborn, P.,… Scales, P. C. (2005). Thriving in adolescence: The voices of youth-serving practitioners, parents, and early and late adolescents. Journal of Early Adolescence, 25, 94–112. doi: 10.1177/0272431604272459 Klimes-Dougan, B., Long, J. D., Lee, C.-Y. S., Ronsaville, D. S., Gold, P. W., & Martinez, P. E. (2010). Continuity and cascade in offspring of bipolar parents: A longitudinal study of externalizing, internalizing, and thought problems. Development and Psychopathology, 22(4), 849–866. doi: 10.1017/S0954579410000507 Kostanjsek, N. (2011). Use of the International Classification of Functioning, Disability and Health (ICF) as a conceptual framework and common language for disability statistics and health information systems. BMC Public Health, 11(Suppl. 4), S3. doi: 10.1186/1471–2458– 11-S4-S3 Kosterman, R., Hawkins, J. D., Spoth, R., Haggerty, K. P., & Zhu, K. (1997). Effects of a preventive parent-training intervention on observed family interactions: proximal outcomes from Preparing for the Drug-Free Years. Journal of Community Psychology, 25, 337–352. doi: 10.1002/(SICI)1520–6629(199707)25:4 Kouros, C. D., Cummings, E. M., & Davies, P. T. (2010). Early trajectories of interparental conflict and externalizing problems as predictors of social competence in preadolescence. Development and Psychopathology, 22(3), 527–537. doi: 10.1017/S0954579410000258 Kraemer, H. C., Kazdin, A., Offord, D. R., Kessler, R. C., Jensen, P. S., & Kupfer, D. (1997). Coming to terms with the terms of risk. Archives of General Psychiatry, 54, 337–343. doi: 10.1001/archpsyc.1997.01830160065009 Kraemer, H. C., Stice, E., Kazdin, A., Offord, D., & Kupfer, D. (2001). How do risk factors work together? Mediators, moderators, and independent, overlapping, and proxy risk factors. American Journal of Psychiatry, 158(6), 848–856. doi: 10.1176/appi.ajp.158.6.848 Kurtines, W. M., Ferrer-Wreder, L., Berman, S. L., Lorente, C. C., Briones, E., Montgomery, M. J.,… Arrufat, O. (2008). Promoting positive youth development: The Miami Youth Development Project (YDP). Journal of Adolescent Research, 23(3), 256–267. doi: 10.1177/0743558408314375 Kutash, K., Lynn, N., & Burns, B. J. (2008). Child and adolescent measures of functional status. In A. J. Rush & D. Blacker (Eds.), Handbook of psychiatric measures (2nd ed., pp.

343–371). Arlington, VA: American Psychiatric Publishing. Lahey, B. B. (2004). Commentary: Role of temperament in developmental models of psychopathology. Journal of Clinical Child and Adolescent Psychology, 33(1), 88–93. doi: 10.1207/s15374424jccp3301_9 Lansford, J. E., Malone, P. S., Dodge, K. A., Pettit, G. S., & Bates, J. E. (2010). Developmental cascades of peer rejection, social information processing biases, and aggression during middle childhood. Development and Psychopathology, 22(3), 593–602. doi: 10.1017/S0954579410000301 Leclair, N., Leclair, S., & Brigham, C. R. (2009). The medical model of impairment. In S. Goldstein & J. A. Naglieri (Eds.), Assessing impairment: From theory to practice (pp. 59– 75). New York, NY: Springer Science. Lehman, A. F. (2009). Disentangle diagnosis and disability. World Psychiatry, 8(2), 89. Lenzenweger, M. F. (2010). A source, a cascade, a schizoid: A heuristic proposal from the Longitudinal Study of Personality Disorders. Development and Psychopathology, 22(4), 867– 881. doi: 10.1017/S0954579410000519 Lerner, J. V., Bowers, E. P., Minor, K., Boyd, M. J., Mueller, M. K.,… Lerner, R. M. (2012). Positive youth development: Processes, philosophies, and programs. In R. M. Lerner, M. A. Easterbrooks, & J. Mistry (Eds.), Handbook of psychology, Vol. 6: Developmental psychology (2nd ed., pp. 365–392). Hoboken, NJ: Wiley. Lerner, R. M. (2002). Concepts and theories of human development (3rd ed.). Mahwah, NJ: Erlbaum. Lerner, R. M., & Benson, P. L. (Eds.). (2003). Developmental assets and asset-building communities: Implications for research, policy, and practice. New York, NY: Kluwer Academic/Plenum. Lerner, R. M., Lerner, J. V., Almerigi, J., Theokas, C., Phelps, E., Gestsdottir, S.,… von Eye, A. (2005). Positive youth development, participation in community youth development programs, and community contributions of fifth grade adolescents: Findings from the first wave of the 4-H Study of Positive Youth Development. Journal of Early Adolescence, 25(1), 17–71. doi: 10.1177/0272431604272461 Lerner, R. M., Lerner, J. V., Lewin-Bizan, S., Bowers, E. P., Boyd, M., Mueller, M., Schmid, K., & Napolitano, C. (2011). Positive youth development: Processes, programs, and problematics. Journal of Youth Development, 6(3), 40–64. Lerner, R. M., Schwartz, S. J., & Phelps, E. (2009). Problematics of time and timing in the longitudinal study of human development: Theoretical and methodological issues. Human Development, 52, 44–68. doi: 10.1159/000189215

Lerner, R. M., Von Eye, A., Lerner, J. V., & Lewin-Bizan, S. (2009). Exploring the foundations and functions of adolescent thriving within the 4-H Study of Positive Youth Development: A view of the issues. Journal of Applied Developmental Psychology, 30(5), 567–570. doi: 10.1016/j.appdev.2009.07.002 Lewandowski, L. J., Lovett, B. J., & Gordon, M. (2009). Measurement of symptom severity and impairment. In S. Goldstein & J. A. Naglieri (Eds.), Assessing impairment: From theory to practice (pp. 5–14). New York, NY: Springer Science. Lewin-Bizan, S., Bowers, E. P., & Lerner, R. M. (2010). One good thing leads to another: Cascades of positive youth development among American adolescents. Development and Psychopathology, 22(4), 759–770. doi: 10.1017/S0954579410000441 Lewis, K. M., Bavarian, N., Snyder, F. J., Acock, A., Day, J., DuBois, D. L.,… Flay, B. R. (2012). Direct and mediated effects of a social-emotional and character development program on adolescent substance use. International Journal of Emotional Education, 4, 56–78. Little, T. D., Preacher, K. J., Selig, J. P., & Card, N. A. (2007). New developments in latent variable panel analyses of longitudinal data. International Journal of Behavioral Development, 31(4), 357–365. doi: 10.1177/0165025407077757 Lochman, J. E., & van den Steenhoven, A. (2002). Family-based approaches to substance abuse prevention. Journal of Primary Prevention, 23(1), 49–114. doi: 10.1023/A:1016591216363 Loevinger, J. (1966). The meaning and measurement of ego development. American Psychologist, 21(3), 195–206. doi: 10.1037/h0023376 Losel, F., & Beelman, A. (2003). Effects of child skills training in preventing antisocial behavior: A systematic review of randomized evaluations. Annals of the American Academy of Political and Social Science, 587, 84–109. doi: 10.1177/0002716202250793 Lynne-Landsman, S. D., Bradshaw, C. P., & Ialongo, N. S. (2010). Testing a developmental cascade model of adolescent substance use trajectories and young adult adjustment. Development and Psychopathology, 22(4), 933–948. doi: 10.1017/S0954579410000556 MacKinnon, D. P. (2008). Introduction to statistical mediation analysis. Mahwah, NJ: Erlbaum. Mahoney, J. L., & Bergman, L. R. (2002). Conceptual and methodological considerations in a developmental approach to the study of positive adaptation. Journal of Applied Developmental Psychology, 23(2), 195–217. doi: 10.1016/s0193–3973(02)00104–1 Martin, M. J., Conger, R. D., Schofield, T. J., Dogan, S. J., Widaman, K. F., Donnellan, M. B., & Neppl, T. K. (2010). Evaluation of the interactionist model of socioeconomic status and problem behavior: A developmental cascade across generations. Development and Psychopathology, 22(3), 695–713. doi: 10.1017/S0954579410000374

Masten, A. S. (2001). Ordinary magic: Resilience processes in development. American Psychologist, 56(3), 227–238. doi: 10.1037/0003–066X.56.3.227 Masten, A. S. (2006). Developmental psychopathology: Pathways to the future. International Journal of Behavioral Development, 30(1), 47–54. doi: 10.1177/0165025406059974 Masten, A. S. (2014). Invited commentary: Resilience and positive youth development frameworks in developmental science. Journal of Youth and Adolescence, 43, 1018–1024. doi: 10.1007/s10964–014–0118–7 Masten, A. S., Burt, K. B., & Coatsworth, J. D. (2006). Competence and psychopathology in development. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology (2nd ed., Vol. 3, pp. 696–738). New York, NY: Wiley. Masten, A. S., & Cicchetti, D. (2010). Developmental cascades. Development and Psychopathology, 22(3), 491–495. doi: 10.1017/S0954579410000222 Masten, A. S., & Coatsworth, J. D. (1995). Competence, resilience, and psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology (2nd ed., Vol. 2, pp. 715– 752). New York, NY: Wiley. Masten, A. S., & Coatsworth, J. D. (1998). The development of competence in favorable and unfavorable environments: Lessons from research on successful children. American Psychologist, 53(2), 205–220. doi: 10.1037/0003–066x.53.2.205 Masten, A. S., Desjardins, C. D., McCormick, C. M., Kuo, S. I.-C., & Long, J. D. (2010). The significance of childhood competence and problems for adult success in work: A developmental cascade analysis. Development and Psychopathology, 22(3), 679–694. doi: 10.1017/S0954579410000362 Masten, A. S., Morison, P., Pelligrini, D., & Tellegen, A. (1990). Competence under stress: Risk and protective factors. In J. Rolf, A. S. Masten, D. Cicchetti, K. Nuechterlein, & S. Weintraub (Eds.), Risk and protective factors in the development of psychopathology (pp. 236–256). New York, NY: Cambridge University Press. Masten, A. S., Roisman, G. I., Long, J. D., Burt, K. B., Obradović, J., Riley, J. R.,… Tellegen, A. (2005). Developmental cascades: Linking academic achievement and externalizing and internalizing symptoms over 20 years. Developmental Psychology, 41(5), 733–746. doi: 10.1037/0012–1649.41.5.733 Maxwell, S. E., & Cole, D. A. (2007). Bias in cross-sectional analyses of longitudinal mediation. Psychological Methods, 12(1), 23–44. doi: 10.1037/1082–989X.12.1.23 McArdle, J. J. (2009). Latent variable modeling of differences and changes with longitudinal data. Annual Review of Psychology, 60(1), 577–605. doi: 10.1146/annurev.psych.60.110707.163612

McCarty, C. A., Mason, W. A., Kosterman, R., Hawkins, J. D., Lengua, L. J., & McCauley, E. (2008). Adolescent school failure predicts later depression among girls. Journal of Adolescent Health, 43(2), 180–187. doi: 10.1016/j.jadohealth.2008.01.023 McClain, D. B., Wolchik, S. A., Winslow, E., Tein, J.-Y., Sandler, I. N., & Millsap, R. E. (2010). Developmental cascade effects of the New Beginnings Program on adolescent adaptation outcomes. Development and Psychopathology, 22(4), 771–784. doi: 10.1017/S0954579410000453 McCormick, C. M., Kuo, S. I.-C., Masten, A. S., Berg, C. A., Smith, J., & Antonucci, T. C. (2011). Developmental tasks across the life-span. In K. L. Fingerman (Ed.), Handbook of lifespan development (pp. 117–140). New York, NY: Springer. Meehl, P. E. (1967). Theory-testing in psychology and physics: A methodological paradox. Philosophy of Science, 34, 103–115. doi: 10.1086/288135 Moilanen, K. L., Shaw, D. S., & Maxwell, K. L. (2010). Developmental cascades: Externalizing, internalizing, and academic competence from middle childhood to early adolescence. Development and Psychopathology, 22(3), 635–653. doi: 10.1017/S0954579410000337 Moore, K. A., & Lippman, L. (Eds.). (2005). What do children need to flourish? Conceptualizing and measuring indicators of positive development. New York, NY: Springer. Motti-Stefanidi, F. (2015). Identity development in the context of the risk and resilience framework. In M. Syed & K. McLean (Eds.), Oxford handbook of identity development (pp. 472–489). New York, NY: Oxford University Press. Mueller, M. K., Phelps, E., Bowers, E. P., Agans, J., Urban, J. B., & Lerner, R. M. (2011). Youth development program participation and intentional self-regulation skills: Contextual and individual bases of pathways to positive youth development. Journal of Adolescence, 34(6), 1115–1126. doi: 10.1016/j.adolescence.2011.07.010 Murray-Close, D., Hoza, B., Hinshaw, S. P., Arnold, L. E., Swanson, J., Jensen, P. S.,… Wells, K. (2010). Developmental processes in peer problems of children with attentiondeficit/hyperactivity disorder in the Multimodal Treatment Study of Children with ADHD: Developmental cascades and vicious cycles. Development and Psychopathology, 22(4), 785– 802. doi: 10.1017/S0954579410000465 Nagin, D., & Tremblay, R. E. (1999). Trajectories of boys' physical aggression, opposition, and hyperactivity on the path to physically violent and nonviolent juvenile delinquency. Child Development, 70(5), 1181–1196. doi: 10.1111/1467–8624.00086 Narrow, W. E., Kuhl, E. A., & Regier, D. A. (2009). DSM-V perspectives on disentangling disability from clinical significance. World Psychiatry, 8(2), 88.

National Institute of Mental Health. (2014). NIMH Research Domain Criteria (RDoC). Retrieved July 20, 2015, from http://www.nimh.nih.gov/research-priorities/rdoc/index.shtml National Research Council & Institute of Medicine (2009). Preventing mental, emotional, and behavioral disorders among young people: Progress and possibilities. (M. E. O'Connell, T. Boat, & K. E. Warner, Eds.) Washington, DC: National Academics Press. Nocentini, A., Calamai, G., & Menesini, E. (2012). Codevelopment of delinquent and depressive symptoms across adolescence: Time-invariant and time-varying effects of school and social failure. Journal of Clinical Child & Adolescent Psychology, 41(6), 746–759. doi: 10.1080/15374416.2012.728155 Oesterle, S., Hawkins, J. D., Fagan, A. A., Abbott, R. D., & Catalano, R. F. (2010). Testing the universality of the effects of the Communities That Care prevention system for preventing adolescent drug use and delinquency. Prevention Science, 11(4), 411–423. doi: 10.1007/s11121-010-0178-1 Obradović, J., Burt, K. B., & Masten, A. S. (2010). Testing a dual cascade model linking competence and symptoms over 20 years from childhood to adulthood. Journal of Clinical Child & Adolescent Psychology, 39(1), 90–102. doi: 10.1080/15374410903401120 Obradović, J., & Hipwell, A. (2010). Psychopathology and social competence during the transition to adolescence: The role of family adversity and pubertal development. Development and Psychopathology, 22(3), 621–634. doi: 10.1017/S0954579410000325 Obradović, J., Shaffer, A. E., & Masten, A. S. (2012). Adversity and risk in developmental psychopathology: Progress and future directions. In L. C. Mayes & M. Lewis (Eds.), The Cambridge Handbook of Environment in Human Development (pp. 35–37). New York, NY: Cambridge University Press. doi: 10.1017/cbo9781139016827.004 Ogden, T., & Hagen, K. A. (2008). Treatment effectiveness of Parent Management Training in Norway: A randomized controlled trial of children with conduct problems. Journal of Consulting and Clinical Psychology, 76(4), 607–621. doi: 10.1037/0022–006x.76.4.607 Overton, W. F. (2013). A new paradigm for developmental science: Relationism and relational-developmental systems. Applied Developmental Science, 17(2), 94–107. doi: 10.1080/10888691.2013.778717 Pantin, H., Prado, G., Lopez, B., Huang, S., Tapia, M. I., Schwartz, S. J.,… Branchini, J. (2009). A randomized controlled trial of Familias Unidas for Hispanic adolescents with behavior problems. Psychosomatic Medicine, 71, 987–995. doi: 10.1097/psy.0b013e3181bb2913 Pardini, D. A., Lochman, J. E., & Frick, P. J. (2003). Callous/unemotional traits and socialcognitive processes in adjudicated youths. Journal of the American Academy of Child & Adolescent Psychiatry, 42(3), 364–371. doi: 10.1097/00004583–200303000–00018

Patterson, G. R., DeBaryshe, B. D., & Ramsey, E. (1989). A developmental perspective on antisocial behavior. American Psychologist, 44(2), 329–335. doi: 10.1037/0003– 066X.44.2.329 Patterson, G. R., Forgatch, M. S., & DeGarmo, D. S. (2010). Cascading effects following intervention. Development and Psychopathology, 22(4), 949–970. doi: 10.1017/S0954579410000568 Patterson, G. R., & Stoolmiller, M. (1991). Replications of a dual failure model for boys' depressed mood. Journal of Consulting and Clinical Psychology, 59, 491–498. doi: 10.1037//0022–006x.59.4.491 Payton, J., Weissberg, R. P., Durlak, J. A., Dymnicki, A. B., Taylor, R. D., Schellinger, K. B., & Payton, M. (2008). The positive impact of social and emotional learning for kindergarten to eighth-grade students: Findings from three scientific reviews. Chicago, IL: Collaborative for Academic, Social, and Emotional Learning. Pearl, J. (2000). Causality: Models, reasoning, and inference. New York, NY: Cambridge University Press. Pearl, J. (2012). The causal mediation formula—A guide to the assessment of pathways and mechanisms. Prevention Science, 13(4), 426–436. doi: 10.1007/s11121–011–0270–1 Pearlin, L. I., Schieman, S., Fazio, E. M., & Meersman, S. C. (2005). Stress, health, and the life course: Some conceptual perspectives. Journal of Health and Social Behavior, 46(2), 205–219. doi: 10.1177/002214650504600206 Pemberton, C. K., Neiderhiser, J. M., Leve, L. D., Natsuaki, M. N., Shaw, D. S., Reiss, D., & Ge, X. (2010). Influence of parental depressive symptoms on adopted toddler behaviors: An emerging developmental cascade of genetic and environmental effects. Development and Psychopathology, 22(4), 803–818. doi: 10.1017/S0954579410000477 Petersen, I. T., Bates, J. E., D'Onofrio, B. M., Coyne, C. A., Lansford, J. E., Dodge, K. A.,… Van Hulle, C. A. (2013). Language ability predicts the development of behavior problems in children. Journal of Abnormal Psychology, 122(2), 542–557. doi: 10.1037/a0031963 Peterson, B. (2013). Editorial: From correlations to causation: The value of preventive interventions in studying pathogenic mechanisms in childhood psychiatric disorders. Journal of Child Psychology and Psychiatry, 54(8), 813–815. doi: 10.1111/jcpp.12122 Phelps, E., Balsano, A., Fay, K., Peltz, J., Zimmerman, S., Lerner, R., M., & Lerner, J. V. (2007). Nuances in early adolescent development trajectories of positive and of problematic/risk behaviors: Findings from the 4-H Study of Positive Youth Development. Child and Adolescent Clinics of North America, 16(2), 473–496. doi: 10.1016/j.chc.2006.11.006 Pianta, R. C., & Hamre, B. K. (2009). Conceptualization, measurement, and improvement of

classroom processes: Standardized observation can leverage capacity. Educational Researcher, 38, 109–119. doi: 10.3102/0013189x09332374 Pickles, A., & Angold, A. (2003). Natural categories or fundamental dimensions: On carving nature at the joints and the rearticulation of psychopathology. Development and Psychopathology, 15(3), 529–551. doi: 10.1017/S0954579403000282 Pickles, A., & Hill, J. (2006). Developmental pathways. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology (2nd ed., Vol. 1). Hoboken, NJ: Wiley. Pittman, K., Irby, M., Tolman, J., Yohalem, N., & Ferber, T. (2003). Preventing problems, promoting development, encouraging engagement: Competing priorities or inseparable goals? Washington, DC: Forum for Youth Investment, Impact Strategies. Prado, G., Pantin, H., Briones, E., Schwartz, S. J., Feaster, D., Huang, S.,… Szapocznik, J. (2007). A randomized controlled trial of a parent centered intervention in preventing substance use and HIV risk behaviors in Hispanic adolescents. Journal of Consulting and Clinical Psychology, 75(6), 914–926. doi: 10.1037/0022–006x.75.6.914 Preacher, K. J., Wichman, A. L., MacCallum, R. C., & Briggs, N. E. (2008). Latent growth curve modeling. Los Angeles, CA: Sage. Price, R. H., Cowen, E. L., Lorina, R. P., & Ramos-McKay, J. (Eds.). (1988). 14 ounces of prevention: A casebook for practitioners. Washington, DC: American Psychological Association. Rapee, R. M., Bőgels, S. M., van der Sluis, C. M., Craske, M. G., & Ollendick, T. (2012). Annual research review: Conceptualising functional impairment in children and adolescents. Journal of Child Psychology and Psychiatry, 53(5), 454–468. doi: 10.1111/j.1469– 7610.2011.02479.x Reyna, V. F., & Farley, F. (2006). Risk and rationality in adolescent decision making: Implications for theory, practice, and public policy. Psychological Science in the Public Interest, 7, 1–49. doi: 10.1111/j.1529–1006.2006.00026.x Riggs, N. R., Greenberg, M. T., Kusché, C. A., & Pentz, M. A. (2006). The mediational role of neurocognition in the behavioral outcomes of a social-emotional prevention program in elementary school students: Effects of the PATHS curriculum. Prevention Science, 7(1), 91– 102. doi: 10.1007/s11121–005–0022–1 Roberts, S., & Pashler, H. (2000). How persuasive is a good fit? A comment on theory testing. Psychological Review, 107(2), 358–367. doi: 10.1037/0033–295X.107.2.358 Robbins, R. N., & Bryan, A. (2004). Relationships between future orientation, impulsive sensation seeking, and risk behavior among adjudicated adolescents. Journal of Adolescent Research, 19, 428–445. doi: 10.1177/0743558403258860

Rodgers, J. L. (2010). The epistemology of mathematical and statistical modeling: A quiet methodological revolution. American Psychologist, 65(1), 1–12. doi: 10.1037/a0018326 Roeser, R. W., Eccles, J. S., & Sameroff, A. J. (2000). School as a context of early adolescents' academic and social-emotional development: A summary of research findings. Elementary School Journal, 100, 443–471. doi: 10.1086/499650 Rogosch, F. A., Oshri, A., & Cicchetti, D. (2010). From child maltreatment to adolescent cannabis abuse and dependence: A developmental cascade model. Development and Psychopathology, 22(4), 883–897. doi: 10.1017/S0954579410000520 Roisman, G. I., Booth-Laforce, C., Belsky, J., Burt, K. B., & Groh, A. M. (2013). Moleculargenetic correlates of infant attachment: A cautionary tale. Attachment and Human Development, 1–23. doi: 10.1080/14616734.2013.768790 Roisman, G. I., Newman, D. A., Fraley, R. C., Haltigan, J. D., Groh, A. M., & Haydon, K. C. (2012). Distinguishing differential susceptibility from diathesis–stress: Recommendations for evaluating interaction effects. Development and Psychopathology, 24(2), 389–409. doi: 10.1017/S0954579412000065 Rutter, M. (1979). Protective factors in children's responses to stress and disadvantage. Annals of the Academy of Medicine, Singapore, 8, 324–338. Rutter, M., & Sroufe, L. A. (2000). Developmental psychopathology: Concepts and challenges. Development and Psychopathology, 12(3), 265–296. doi: 10.1017/s0954579400003023 Sameroff, A. J. (2000). Developmental systems and psychopathology. Development and Psychopathology, 12(3), 297–312. doi: 10.1017/s0954579400003035 Sameroff, A. J. (2009). Designs for translational research. In A. J. Sameroff (Ed.), The transactional model of development: How children and contexts shape each other (pp. 23– 32). Washington, DC: American Psychological Association. Sameroff, A. J. (2010). A unified theory of development: A dialectic integration of nature and nurture. Child Development, 81(1), 6–22. doi: 10.1111/j.1467–8624.2009.01378.x Sanders, M. R., Markie-Dadds, C., Tully, L. A., & Bor, W. (2000). The Triple P-Positive Parenting Program: A comparison of enhanced, standard, and self-directed behavioral family intervention for parents of children with early onset conduct problems. Journal of Consulting and Clinical Psychology, 68(4), 624–640. doi: 10.1037/0022–006x.68.4.624 Sandler, I. N., Ayers, T. S., Tein, J. Y., Wolchik, S. A., Millsap, R., Khoo, S. T.,… Coxe, S. (2010). Six-year follow-up of a preventive intervention for parentally-bereaved youth: A randomized controlled trial. Archives of Pediatric and Adolescent Medicine, 164, 907–914. doi: 10.1001/archpediatrics.2010.173 Sandler, I. N., Schoenfelder, E. N., Wolchik, S. A., & MacKinnon, D. P. (2011). Long-term

impact of prevention programs to promote effective parenting: Lasting effects but uncertain processes. Annual Review of Psychology, 62, 299–329. doi: 10.1146/annurev.psych.121208.131619 Sanford, M. N., Offord, D. R., Boyle, M. H., Peace, A., & Racine, Y. A. (1992). Ontario Child Health Study: Social and school impairments in children aged 6 to 16 years. Journal of the American Academy of Child & Adolescent Psychiatry, 31(1), 60–67. doi: 10.1097/00004583–199201000–00010 Sanislow, C. A., Pine, D. S., Quinn, K. J., Kozak, M. J., Garvey, M. A., Heinssen, R. K.,… Cuthbert, B. N. (2010). Developing constructs for psychopathology research: Research domain criteria. Journal of Abnormal Psychology, 119(4), 631–639. doi: 10.1037/a0020909 Schaps, E., Battistich, V., & Solomon, D. (2004). Community in school as key to student growth: Findings from the Child Development Project. In J. E. Zins, R. P. Weissberg, M. C. Wang, & H. J. Walberg (Eds.), Building academic success on social and emotional learning: What does the research say? (pp. 189–205). New York, NY: Teachers College Press. Scheier, L. M., Botvin, G. J., Diaz, T., & Griffin, K. W. (1999). Social skills, competence, and drug refusal efficacy as predictors of adolescent alcohol use. Journal of Drug Education, 29, 251–278. doi: 10.2190/m3ct-wwjm-5jaq-wp15 Schulenberg, J. E., & Maslowsky, J. (2009). Taking substance use and development seriously: Developmentally distal and proximal influences on adolescent drug use. Monographs of the Society for Research in Child Development, 74(3), 121–130. doi: 10.1111/j.1540– 5834.2009.00544.x Schulenberg, J. E., Sameroff, A. J., & Cicchetti, D. (2004). The transition to adulthood as a critical juncture in the course of psychopathology and mental health. Development and Psychopathology, 16(4), 799–806. doi: 10.1017/S0954579404040015 Schwartz, S. J., Pantin, H., Coatsworth, J. D., & Szapocznik, J. (2007). Addressing the challenges and opportunities of today's youth: Toward an integrative model and its implications for research and intervention. Journal of Primary Prevention, 28, 117–144. doi: 10.1007/s10935–007–0084-x Scorza, P., Stevenson, A., Canino, G., Mushashi, C., Kanyanganzi, F., Munyanah, M., & Betancourt, T. (2013). Validation of the “World Health Organization Disability Assessment Schedule for Children, WHODAS-Child” in Rwanda. PloS One, 8(3), e57725. doi: 10.1371/journal.pone.0057725 Selig, J. P., & Preacher, K. J. (2009). Mediation models for longitudinal data in developmental research. Research in Human Development, 6(2–3), 144–164. doi: 10.1080/15427600902911247 Selig, J. P., Preacher, K. J., & Little, T. D. (2012). Modeling time-dependent association in

longitudinal data: A lag as moderator approach. Multivariate Behavioral Research, 47(5), 697–716. doi: 10.1080/00273171.2012.715557 Shaffer, D., Gould, M. S., Brasic, J., Ambrosini, P., Fisher, P., Bird, H., & Aluwahlia, S. (1983). A Children's Global Assessment Scale (CGAS). Archives of General Psychiatry, 40(11), 1228–1231. doi: 10.1001/archpsyc.1983.01790100074010 Sigal, A. B., Wolchik. S. A., Tien, J. Y., & Sandler, I. N. (2012). Enhancing youth outcomes following parental divorce: a longitudinal study of the effects of the New Beginnings Program on educational and occupational goals. Journal of Clinical Child and Adolescent Psychology, 41, 150–165. doi: 10.1080/15374416.2012.651992 Sigmarsdóttir, M., DeGarmo, D. S., Forgatch, M. S., & Guðmundsdóttir, E. V. (2013). Treatment effectiveness of PMTO for children's behavior problems in Iceland: Assessing parenting practices in a randomized controlled trial. Scandinavian Journal of Psychology, 54(6), 468–476. doi: 10.1111/sjop.12078 Simpson, A. R. (2001). Raising Teens: A Synthesis of Research and a Foundation for Action. Boston: Center for Health Communication, Harvard School of Public Health. http://hrweb.mit.edu/worklife/raising-teens/pdfs/raising_teens_report.pdf Simpson, A. R., & Roehlkepartain, J. L. (2003). Asset building in parenting practices and family life. In R. M. Lerner & P. L. Benson (Eds.), Developmental assets and asset-building communities (pp. 157–193). New York: Springer. doi: 10.1007/978-1-4615-0091-9_7 Slavich, G. M., & Cole, S. W. (2013). The emerging field of human social genomics. Clinical Psychological Science, 1(3), 331–348. doi: 10.1177/2167702613478594 Smith, G. N., Ehmann, T. S., Flynn, S. W., MacEwan, G. W., Tee, K., Kopala, L. C.,… Honer, W. G. (2011). The assessment of symptom severity and functional impairment with DSM-IV Axis V. Psychiatric Services, 62(4), 411–417. doi: 10.1176/appi.ps.62.4.411 Snyder, F. J., Acock, A. C., Vuchinich, S., Beets, M. W., Washburn, I. J., & Flay, B. R. (2013). Preventing negative behaviors among elementary-school students through enhancing students' social-emotional and character development. American Journal of Health Promotion, 28(1), 50–58. doi: 10.4278/ajhp.120419-QUAN-207 Snyder, F. J., Flay, B. R., Vuchinich, S., Acock, A., Washburn, I. J., Beets, M., & Li, K. (2010). Impact of a social-emotional and character development program on school-level indicators of academic achievement, absenteeism, and disciplinary outcomes: A matched-pair, clusterrandomized, controlled trial. Journal of Research on Educational Effectiveness, 3, 26–55. doi: 10.1080/19345740903353436 Snyder, F. J., Vuchinich, S., Acock, A., Washburn, I. J., & Flay, B. R. (2012). Improving elementary school quality through the use of a social-emotional and character development program: A matched-pair, cluster-randomized, controlled trial in Hawai‘i. Journal of School

Health, 82, 11–20. doi: 10.1111/j.1746–1561.2011.00662.x Soper, A. C., Wolchik, S. A., Tein, J-.Y., & Sandler, I. N. (2010). Mediation of a preventive intervention's six-year effects program effects on health risk behaviors. Psychology of Addictive Behaviors, 24, 300–310. doi: 10.1037/a0019014 Sparrow, S. S., Cicchetti, D. V., & Balla, D. (2005). Vineland Adaptive Behavior Scales (2nd ed.). Circle Pines, MN: AGS. Spilt, J. L., van Lier, P. A. C., Leflot, G., Onghena, P., & Colpin, H. (2014). Children's selfconcept and internalizing problems: The influence of peers and teachers. Child Development, 85(3), 1248–1256. doi: 10.1111/cdev.12181 Spoth, R. L., Kavanagh, K., & Dishion, T. J. (2002). Family-centered preventive intervention science: Toward benefits to larger populations of children, youth, and families. Prevention Science, 3, 145–152. doi: 10.1023/a:1019924615322 Spoth, R., Randall, G. K., & Shin, C. (2008). Increasing school success through partnershipbased family competency training: Experimental study of long-term outcomes. School Psychology Quarterly, 23(1), 70–89. doi: 10.1037/1045–3830.23.1.70 Spoth, R., Shin, C., Guyll, M., Redmond, C., & Azevedo, K. (2006). Universality of effects: An examination of the comparability of long-term family intervention effects on substance use across risk-related subgroups. Prevention Science, 7, 209–224. doi: 10.1007/s11121–006– 0036–3 Sroufe, L. A. (1979). The coherence of individual development: Early care, attachment, and subsequent developmental issues. American Psychologist, 34, 834–841. doi: 10.1037//0003– 066x.34.10.834 Sroufe, L. A. (1997). Psychopathology as an outcome of development. Development and Psychopathology, 9(2), 251–268. doi: 10.1017/S0954579497002046 Sroufe, L. A., Egeland, B., & Kreutzer, T. (1990). The fate of early experience following developmental change: Longitudinal approaches to individual adaptation in childhood. Child Development, 61, 1363–1373. doi: 10.1111/j.1467–8624.1990.tb02867.x Sroufe, L. A., & Rutter, M. (1984). The domain of developmental psychopathology. Child Development, 55(1), 17–29. doi: 10.2307/1129832 Staff, J., Schulenberg, J. E., Maslowsky, J., Bachman, J. G., O'Malley, P. M., Maggs, J. L., & Johnston, L. D. (2010). Substance use changes and social role transitions: Proximal developmental effects on ongoing trajectories from late adolescence through early adulthood. Development and Psychopathology, 22(4), 917–932. doi: 10.1017/S0954579410000544 Sternberg, R. J. (1985). Beyond IQ: A triarchic theory of human intelligence. New York, NY: Cambridge University Press.

Stormshak, E. A., Connell, A., & Dishion, T. J. (2009). An adaptive approach to familycentered intervention in schools: Linking intervention engagement to academic outcomes in middle and high school. Prevention Science, 10, 221–235. doi: 10.1007/s11121–009–0131–3 Stringaris, A., & Goodman, R. (2013). The value of measuring impact alongside symptoms in children and adolescents: A longitudinal assessment in a community sample. Journal of Abnormal Child Psychology, 41(7), 1109–1120. doi: 10.1007/s10802–013–9744-x Sturaro, C., van Lier, P. A. C., Cuijpers, P., & Koot, H. M. (2011). The role of peer relationships in the development of early school-age externalizing problems. Child Development, 82(3), 758–765. doi: 10.1111/j.1467–8624.2010.01532.x Task Panel of Prevention. (1978). Task panel reports submitted to the President's Commission on Mental Health (Vol. 4). Washington, DC: U.S. Government Printing Office. Taylor, T. K., & Biglan, A. (1999). Behavioral parenting skills programs: A review of the literature for clinicians and policy makers. Clinical Child and Family Psychology Review, 2, 169–182. Thornberry, T. P., Freeman-Gallant, A., & Lovegrove, P. J. (2009a). The impact of parental stressors on the intergenerational transmission of antisocial behavior. Journal of Youth and Adolescence, 38(3), 312–322. doi: 10.1007/s10964–008–9337–0 Thornberry, T. P., Freeman-Gallant, A., & Lovegrove, P. J. (2009b). Intergenerational linkages in antisocial behaviour. Criminal Behaviour and Mental Health, 19(2), 80–93. doi: 10.1002/cbm.709 Tobler, N. S., Roona, M. R., Ochshorn, P., Marshall, D. G., Streke, A. V., & Stackpole, K. M. (2000). School-based adolescent drug prevention programs: 1998 meta-analysis. Journal of Primary Prevention, 20(4), 275–336. doi: 10.1023/a:1021314704811 Tolan, P. H. (2002). Family-focused prevention research: “Tough but tender.” In H. A. Liddle, D. A. Santisteban, R. L. Levant, & J. H. Bray (Eds.), Family psychology: Science-based interventions (pp. 197–213). Washington, DC: American Psychological Association. Tolan, P. H., Gorman-Smith, D., & Henry, D. (2004). Supporting families in a high-risk setting: Proximal effects of the SAFE Children Preventive Intervention. Journal of Consulting and Clinical Psychology, 72, 855–869. doi: 10.1037/0022–006x.72.5.855 Tolan, P. H., Gorman-Smith, D., Henry, D., & Schoeny, M. (2009). The benefits of booster interventions: Evidence from a family-focused prevention program. Prevention Science, 10, 287–297. doi: 10.1007/s11121–009–0139–8 Tram, J. M., & Cole, D. A. (2000). Self-perceived competence and the relation between life events and depressive symptoms in adolescence: Mediator or moderator? Journal of Abnormal Psychology, 109, 753–760. doi: 10.1037/0021–843x.109.4.753

Trudeau, L., Spoth, R., Randall, G. K., & Azevedo, K. (2007). Longitudinal effects of a universal family-focused intervention on growth patterns of adolescent internalizing symptoms and polysubstance use: Gender comparisons. Journal of Youth and Adolescence, 36, 725–740. doi: 10.1007/s10964–007–9179–1 Tully, E. C., Iacono, W. G., & McGue, M. (2010). Changes in genetic and environmental influences on the development of nicotine dependence and major depressive disorder from middle adolescence to early adulthood. Development and Psychopathology, 22(4), 831–848. doi: 10.1017/S0954579410000490 United Nations Children's Fund. (2007). Child poverty in perspective: An overview of child well-being in rich countries: Innocenti Report Card 7. Florence, Italy: UNICEF Innocenti Research Centre. Üstün, B., Chatterji, S., Kostanjsek, N., Rehm, J., Kennedy, C., Epping-Jordan, J.,… Pull, C. (2010). Developing the World Health Organization Disability Assessment Schedule 2.0. Bulletin of the World Health Organization, 88(11), 815–823. doi: 10.2471/BLT.09.067231 Üstün, B., & Kennedy, C. (2009). What is “functional impairment”? Disentangling disability from clinical significance. World Psychiatry, 8(2), 82. Vaillancourt, T., Brittain, H. L., McDougall, P., & Duku, E. (2013). Longitudinal links between childhood peer victimization, internalizing and externalizing problems, and academic functioning: Developmental cascades. Journal of Abnormal Child Psychology, 41, 1203– 1215. doi: 10.1007/s10802–013–9781–5 van der Knaap, L. M., Schulze, H.-J., Slot, N. W., & Feij, J. A. (2009). Tasks and skills in youth with problem behaviour. Development of a questionnaire. Vulnerable Children and Youth Studies, 4(3), 264–276. doi: 10.1080/17450120802626346 van IJzendoorn, M. H., & Bakermans-Kranenburg, M. J. (2015). Genetic differential susceptibility on trial: Meta-analytic support from randomized controlled experiments. Development and Psychopathology, 27(1), 151–162. doi: 10.1017/S0954579414001369 van Lier, P. A. C., & Koot, H. M. (2010). Developmental cascades of peer relations and symptoms of externalizing and internalizing problems from kindergarten to fourth-grade elementary school. Development and Psychopathology, 22(3), 569–582. doi: 10.1017/S0954579410000283 van Lier, P. A. C., Vitaro, F., Barker, E. D., Brendgen, M., Tremblay, R. E., & Boivin, M. (2012). Peer victimization, poor academic achievement, and the link between childhood externalizing and internalizing problems. Child Development, 83(5), 1775–1788. doi: 10.1111/j.1467–8624.2012.01802.x Van Ryzin, M. J., & Dishion, T. J. (2013). From antisocial behavior to violence: A model for the amplifying role of coercive joining in adolescent friendships. Journal of Child Psychology

and Psychiatry, 54(6), 661–669. doi: 10.1111/jcpp.12017 Velez, C., Wolchik, S. A., Tein, J. Y., & Sandler, I. N. (2011). Protecting children from the consequences of divorce: A longitudinal study of the effects of parenting on children's coping efforts and coping efficacy. Child Development, 82, 244–257. doi: 10.1111/j.1467– 8624.2010.01553.x Véronneau, M.-H., Vitaro, F., Brendgen, M., Dishion, T. J., & Tremblay, R. E. (2010). Transactional analysis of the reciprocal links between peer experiences and academic achievement from middle childhood to early adolescence. Developmental Psychology, 46(4), 773–790. doi: 10.1037/a0019816 Von Eye, A., & Bergman, L. R. (2003). Research strategies in developmental psychopathology: Dimensional identity and the person-oriented approach. Development and Psychopathology, 15(3), 553–580. doi: 10.1017/S0954579403000294 Wakefield, J. C. (1997). When is development disordered? Developmental psychopathology and the harmful dysfunction analysis of mental disorder. Development and Psychopathology, 9(2), 269–290. doi: 10.1017/S0954579497002058 Wakefield, J. C. (2009). Disability and diagnosis: Should role impairment be eliminated from DSM/ICD diagnostic criteria? World Psychiatry, 8(2), 87. Wakschag, L. S., Choi, S. W., Carter, A. S., Hullsiek, H., Burns, J., McCarthy, K.,… BriggsGowan, M. J. (2012). Defining the developmental parameters of temper loss in early childhood: Implications for developmental psychopathology. Journal of Child Psychology and Psychiatry, 53(11), 1099–1108. doi: 10.1111/j.1469–7610.2012.02595.x Wakschag, L. S., Henry, D. B., Tolan, P. H., Carter, A. S., Burns, J. L., & Briggs-Gowan, M. J. (2012). Putting theory to the test: Modeling a multidimensional, developmentally-based approach to preschool disruptive behavior. Journal of the American Academy of Child and Adolescent Psychiatry, 51(6), 593–604. doi: 10.1016/j.jaac.2012.03.005 Washburn, I. J., Acock, A., Vuchinich, S., Snyder, F., Li, K. K., Ji, P.,… Flay, B. R. (2011). Effects of a social emotional and character development program on the trajectory of behaviors associated with social emotional and character development: Findings from three randomized trials. Prevention Science, 12(3), 314–323. doi: 10.1007/s11121–011–0230–9 Waters, E., & Sroufe, L. A. (1983). Social competence as a developmental construct. Developmental Review, 3, 79–97. doi: 10.1016/0273–2297(83)90010–2 Wechsler, D. (1958). The measurement and appraisal of adult intelligence. Baltimore, MD: Williams & Wilkins. Weissberg, R. P., & Greenberg, M. T. (1998). School and community competence-enhancement and prevention programs. In I. E. Sigel & K. A. Renninger (Eds.), Handbook of child psychology: Vol 4. Child psychology in practice (5th ed., pp. 877–954). Hoboken, NJ: Wiley.

Weissman, M. M. (2009). Functional impairment can have different meanings. World Psychiatry, 8(2), 94. Weisz, J. R., Sandler, I. N., Durlak, J. A. & Anton, B. S. (2005). Promoting and protecting youth mental health through evidence-based prevention and treatment. American Psychologist, 60, 628–648. doi: 10.1037/0003–066x.60.6.628 Wentzel, K. R., & Wigfield, A. (1998). Academic and social motivational influences on students' academic performance. Educational Psychology Review, 10, 155–175. doi: 10.1023/a:1022137619834 White, R. W. (1959). Motivation reconsidered: The concept of competence. Psychological Review, 66, 297–333. doi: 10.1037/h0040934 Whiteford, H. (2009). Clarifying the relationship between symptoms and disability: A challenge with practical implications. World Psychiatry, 8(2), 90. Whiteside, S. P. (2009). Adapting the Sheehan Disability Scale to assess child and parent impairment related to childhood anxiety disorders. Journal of Clinical Child & Adolescent Psychology, 38(5), 721–730. doi: 10.1080/15374410903103551 Wills, T. A., Walker, C., Mendoza, D., & Ainette, M. G. (2006). Behavioral and emotional self-control: Relations to substance use in samples of middle and high school students. Psychology of Addictive Behaviors, 20, 265–278. doi: 10.1037/0893–164x.20.3.265 Wilson, D. S., O'Brien, D. T., & Sesma, A. (2009). Human prosociality from an evolutionary perspective: Variation and correlations at a citywide scale. Evolution and Human Behavior, 30, 190–200. doi: 10.1016/j.evolhumbehav.2008.12.002 Wilson, S. J., & Lipsey, M. W. (2007). School-based interventions for aggressive and disruptive behavior: Update of a meta-analysis. American Journal of Preventive Medicine, 33(Suppl. 2S), 130–143. doi: 10.1016/j.amepre.2007.04.011 Winters, N. C., Collett, B. R., & Myers, K. M. (2005). Ten-year review of rating scales, VII: Scales assessing functional impairment. Journal of the American Academy of Child & Adolescent Psychiatry, 44(4), 309–338. doi: 10.1097/01.chi.0000153230.57344.cd Wolchik, S. A., Sandler, I. N., Jones, S., Gonzales, N., Doyle, K., Winslow, E.,… Braver, S. L. (2009). The New Beginnings Program for divorcing and separating families: Moving from efficacy to effectiveness. Family Court Review, 47(3), 416–435. doi: 10.1111/j.1744– 1617.2009.01265.x Wolchik, S. A., Sandler, I. N., Millsap, R. E., Plummer, B. A., Greene, S. M., Anderson, E. R., … Haine, R. A. (2002). Six-year follow-up of preventive interventions for children of divorce: A randomized controlled trial. Journal of the American Medical Association, 288, 1874–1881. doi: 10.1001/jama.288.15.1874

Wolchik, S. A., Sandler, I. N., Weiss, L., & Winslow, E. (2007). New Beginnings: An empirically-based program to help divorced mothers promote resilience in their children. In J. M. Briesmeister & C. E. Schaefer (Eds.), Handbook of parent training (pp. 25–66). Hoboken, NJ: Wiley. Wolchik, S. A., West, S. G., Sandler, I. N., Tein, J., Coatsworth, D., Lengua, L.,… Griffin, W. A. (2000). An experimental evaluation of theory-based mother and mother–child programs for children of divorce. Journal of Consulting and Clinical Psychology, 68, 843–856. doi: 10.1037//0022–006x.68.5.843 World Health Organization. (2001). International Classification of Functioning, Disability and Health: ICF. Geneva, Switzerland: Author. World Health Organization. (2007). International Classification of Functioning, Disability, and Health for Children and Youth (ICF-CY). Geneva, Switzerland: Author. Yates, T. M., Burt, K. B., & Troy, M. F. (2011). A developmental approach to clinical research, classification, and practice. In D. Cicchetti & G. I. Roisman (Eds.), The origins and organization of adaptation and maladaptation (Vol. 36, pp. 231–282). Hoboken, NJ: Wiley. Yates, T. M., Obradović, J., & Egeland, B. (2010). Transactional relations across contextual strain, parenting quality, and early childhood regulation and adaptation in a high-risk sample. Development and Psychopathology, 22(3), 539–555. doi: 10.1017/S095457941000026X Yong, M., Fleming, C. B., McCarty, C. A., & Catalano, R. F. (2014). Mediators of the associations between externalizing behaviors and internalizing symptoms in late childhood and early adolescence. Journal of Early Adolescence, 34(7), 967–1000. doi: 10.1177/0272431613516827 Zarrett, N., Peltz, J., Fay, K., Carrano, J., Li, Y., & Lerner, R. M. (2007). Sports and youth development programs: Theoretical and practical implications of early adolescent participation in multiple instances of structured out-of-school-time (OST) activity. Journal of Youth Development, 2, 7–20. Zelazo, P. D., & Carlson, S. M. (2012). Hot and cool executive function in childhood and adolescence: Development and plasticity. Child Development Perspectives, 6(4), 354–360. doi: 10.1111/j.1750–8606.2012.00246.x Zhou, Q., Hofer, C., Eisenberg, N., Reiser, M., Spinrad, T. L., & Fabes, R. A. (2007). The developmental trajectories of attention focusing, attentional and behavioral persistence, and externalizing problems during school-age years. Developmental Psychology, 43(2), 369–385. doi: 10.1037/0012–1649.43.2.369 Zhou, Q., Sandler, I. N., Millsap, R. E., Wolchik, S. A., & Dawson-McClure, S. R. (2008). Mother-child relationship quality and effective discipline as mediators of the 6-year effects of the New Beginnings Program for children from divorced families. Journal of Consulting and

Clinical Psychology, 76(4), 579–594. doi: 10.1037/0022–006X.76.4.579 Zins, J. E., Weissberg, R. P., Wang, M. C., & Walberg, H. J. (2004). Building school success through social and emotional learning. New York, NY: Teachers College Press. Zwirs, B., Burger, H., Schulpen, T., Vermulst, A. A., HiraSing, R. A., & Buitelaar, J. (2011). Teacher ratings of children's behavior problems and functional impairment across gender and ethnicity: Construct equivalence of the Strengths and Difficulties Questionnaire. Journal of Cross-Cultural Psychology, 42(3), 466–481. doi: 10.1177/0022022110362752

Chapter 10 The Development of Coping: Implications for Psychopathology and Resilience Melanie J. Zimmer-Gembeck and Ellen A. Skinner It's a characteristic of human nature that the best qualities, called up quickly in a crisis, are very often the hardest to find in prosperous calm. The contours of all our virtues are shaped by adversity. —Gregory David Roberts, Shantaram

GOAL OF THE CHAPTER TRANSACTIONAL PERSPECTIVES: COPING AS INDIVIDUAL DIFFERENCES IN APPRAISAL AND COPING PROCESSES AND RESOURCES Stress, Appraisals, and Coping Associated with Adjustment and Psychopathology Links between Broad Categories of Coping and Psychopathology Do Subjective Appraisals of Stressful Encounters also Play a Role in Psychopathology? Strategies for Emotion Regulation, Coping, and Psychopathology Patterns or Profiles of Coping as Correlates of Psychopathology Transactional Models of the Links between Stress, Coping, and Psychopathology Critique of Individual Differences Research on Coping and Psychopathology NORMATIVE DEVELOPMENTAL PERSPECTIVES: COPING AS A SET OF BASIC ADAPTIVE PROCESSES THAT ARE REORGANIZED WITH AGE Normative Development of Coping during Infancy: Implicit Coping Normative Development of Coping during Early Childhood: Voluntary Coping Normative Development of Coping during Middle Childhood: Reflective Coping Normative Development of Coping during Adolescence: Proactive Coping Conclusion Normative Development of Coping and Developmental Psychopathology DEVELOPMENTAL SYSTEMS PERSPECTIVES: COPING AS PART OF DEVELOPMENTAL CASCADES TOWARD PSYCHOPATHOLOGY AND RESILIENCE Temperament, Differential Pathways of Maladaptive Coping, and Psychopathology Attachment, Differential Pathways of Maladaptive Coping, and Psychopathology

Parenting, Differential Pathways of Maladaptive Coping, and Psychopathology Family Stress, Differential Pathways of Maladaptive Coping, and Psychopathology FUTURE RESEARCH AND TRANSLATION OF RESEARCH INTO ACTION The Role of Coping in Developmental Cascades toward Psychopathology and Resilience Translation of Basic Research on Coping into Action SUMMARY AND CONCLUSION REFERENCES A primary justification for the study of coping is the notion that, when people are faced with adversity, the ways they react to and deal with its challenges can make a material difference to their subsequent development. If they are overwhelmed, they can become more vulnerable to subsequent psychological problems and disorder; if they rise to the challenge, they can become toughened, strengthened, and more resilient to future threats and difficulties. Akin to the concept of host resistance in the study of whether exposure to germs will lead to illness or to immunity, the concept of coping refers not to the assets and liabilities people bring to their dealings with adversity but instead to how people actually interact with the real problems, setbacks, and difficulties they encounter daily, right on the ground. These myriad ways of coping, such as problem solving, negotiation, rumination, accommodation, escape, confrontation, and help seeking, describe specific transactions along the arc of episodic encounters with stress and suggest one set of mechanisms through which adversity can erode individual resources to create long-term liabilities or, alternatively, can help individuals accrue lasting competencies for managing stress. Surprisingly, however, the strands of research that focus on coping have not been well integrated with research on resilience and the development of psychopathology. Despite the potential centrality of coping as a moderator and mediator of exposure to stress, coping is rarely included in programs of study focusing on developmental psychopathology or resilience (cf. Cicchetti & Rogosch, 2009). By the same token, despite the fact that the adversities people face, such as poverty and maltreatment, shape both the array of problems they confront and the resources they can access to deal with them, coping research rarely considers the effects of higher-order contexts (cf. Tolan & Grant, 2009). All of these areas would benefit from a further exploration of their inherent interconnections. The natural overlaps among coping, developmental psychopathology, and resilience have been expanded over the last several decades by a major shift in conceptualizations guiding the study of coping during childhood and adolescence. This shift was initiated in the 1980s with the publication of the book Stress, Coping, and Development in Children (edited by Garmezy & Rutter, 1988) and the seminal article in Psychological Bulletin titled “Coping with Stress during Childhood and Adolescence” (Compas, 1987a), and then it accelerated rapidly during the late 1990s (Eisenberg, Fabes, & Guthrie, 1997; Skinner & Edge, 1998; Wolchik & Sandler, 1997). Since that time, conceptualizations of coping during childhood and adolescence have

branched off from work on coping during adulthood (Aldwin, 2007), after which it was largely patterned, to focus not only on individual differences but also on the development of coping. This “new” way of looking at coping, which takes the concept back to its roots as an adaptive process (Murphy & Moriarity, 1976; White, 1974), has forced a reconsideration of the very definition of coping as well as the meaning of ways of coping, their antecedents and consequences, the role of social partners, and, most important, qualitative shifts with age in how the adaptive processes that comprise coping are organized across infancy, childhood, adolescence, and early adulthood (Skinner & Zimmer-Gembeck, 2007, in press).

Goal of the Chapter The goal of this chapter is to review conceptual and empirical progress in the study of the development of coping and to identify important ways in which this work may be useful to researchers studying the development of psychopathology and resilience. We present our review and ideas in three main sections, each representing a different perspective on coping. An overview of the contributions of each of these perspectives to an understanding of the connections among coping, psychopathology, and resilience is provided in Table 10.1. First we focus on coping as a transactional process, basically as constituting individual differences in appraisals and ways of dealing with stressful demands and the emotions they generate. Here we review what is known about the links between different ways of coping, appraisals, and coping resources, on one hand, and different forms of psychopathology and adaptive functioning, on the other hand. We primarily draw from research on children and adolescents, but research with adults is also reviewed when it is important for identifying future directions for research with youth. We also provide conceptual and methodological critiques of this work to date. Table 10.1 Three Perspectives on How the Study of the Development of Coping Can Contribute to Research on Developmental Psychopathology and Resilience Transactional Perspectives: Coping as Individual Differences in Stress Reactivity and Responses Coping is a moderator. Ways of coping represent families of adaptive and maladaptive responses to stress and so fundamentally buffer or exacerbate the effects of stress and adversity on psychopathology and resilience. Coping is a mediator. Adaptive and maladaptive coping are parts of the pathways through which exposure to stress contributes to psychopathology and resilience. Individual ways of coping are part of a profile of reactions and responses during the course of stressful encounters. Coping can supplement the study of individual ways of coping implicated in psychopathology (such as rumination or social isolation) by considering the repertoire, combination, or sequence of ways of coping that children and adolescents enact over the arc of multiple coping episodes. Coping is a critical mechanism through which a variety of assets and liabilities, such as self-

efficacy, pessimism, and social support, which have long been implicated in psychopathology and resilience, exert their effects. Normative Developmental Perspectives: Coping as a Fundamental Adaptive Process Developmentally friendly conceptualizations, which define coping as regulation under stress, integrate research on the development of both stress reactivity and the many kinds of regulation (emotional, attentional, motivational, behavioral, etc.) that are activated by stressful encounters. Because core families of coping represent fundamental adaptive and maladaptive processes, they are tightly linked to other subsystems that serve to detect and deal with threats and danger, such as processes of temperament, attachment, mastery, and self-determination. Coping influences everyday resilience and marks a site of developmental potential. Coping shapes how children and youth bounce back from daily stressors. Such episodes can be opportunities for the development of regulatory capacities and coping efficacy, if stressors are manageable, personal and interpersonal resources are sufficient, and parents (and other adults) help children channel setbacks and failures adaptively—by learning and growing from them. Developmental Systems Perspectives: Coping as Part of Developmental Cascades The stresses of adversity can undermine the healthy development of coping—based on its effects on the neurophysiological, psychological, and social underpinnings of coping. The development of adaptive coping requires years of deep developmentally attuned interpersonal support for dealing with just-manageable demands. Stressful overarching social conditions, such as poverty, oppression, discrimination, harsh families and parenting, maltreatment, and neglect, pose serious risks to the healthy development of coping. Coping is reciprocally related to psychopathology and resilience. Psychopathology interferes with constructive coping and triggers maladaptive stress reactions. Hence, coping is a key marker and player in the developmental cascades from which psychopathology emerges. Coping can be a powerful intervention lever in preventing deleterious short- and long-term consequences of stress, risk, and adversity. Upstream interventions that focus on strengthening relationships and promoting core coping resources and coping efficacy should have the biggest impact on the construction of appraisals and coping processes that foster resilience. Then we describe coping as a fundamental human adaptive process that involves the regulation of multiple subsystems (e.g., emotion and attention) that are activated by stress and that also shows regular age-graded developments in how such regulation is accomplished. We explain how a developmental perspective requires a reorientation of the study of coping, including a developmentally friendly definition of coping itself. We review studies of normative age differences and age changes in adaptive and maladaptive ways of coping and knit together research from within and outside the coping area to create a picture of the neurophysiological underpinnings and qualitative shifts in coping as it develops from birth to emerging adulthood. This section highlights the important roles that close relationships with caring adults play in the healthy development of coping and explains how stressful encounters

can provide opportunities for the development of coping and regulatory capacities, resources, and efficacy, if demands are manageable, interpersonal supports are sufficient, and parents (and other adults) help children channel setbacks and failures adaptively—by learning and growing from them. We describe how these basic building blocks all work together to contribute to the normative development of appraisals, coping, and personal and social coping resources in order to show how they inherently provide a platform or foundation for more complex theories and research on the development of psychopathology or resilience. The section titled “Developmental Systems Perspectives” brings together work on individual differences and normative development to consider coping from a multilevel dynamic developmental systems perspective. According to this approach, coping is always part of developmental cascades that contribute to resilience or psychopathology. Profiles of adaptive and maladaptive coping can be considered both markers and mechanisms for cascades leading in a multitude of directions, and ways of coping are important processes that shuttle individuals back and forth between different pathways. We draw from work on temperament, attachment, and family stress to identify examples of underlying biological and overarching contextual risk and protective factors that shape developmental transitions and dynamics and so likely play a role in developmental cascades that mark and contribute to psychopathology and resilience. We end the chapter with some suggestions about how future research can productively combine work on the development of coping with work on the development of psychopathology and resilience.

Transactional Perspectives: Coping as Individual Differences in Appraisal and Coping Processes and Resources Early work on coping during childhood and adolescence, inspired by research with adults, focused on coping as a transactional process that unfolds in several recursive steps (Folkman & Moskowitz, 2004; Lazarus & Folkman, 1984). According to this perspective, as pictured in Figure 10.1, coping transactions are initiated by encounters with stress, defined as internal and external events that individuals appraise as important to their well-being and as taxing or exceeding their resources (Lazarus & Folkman, 1986, p. 63). Cognitive appraisals, focusing on the extent to which the stressor is personally relevant and amenable to personal control, result in views of the encounter as constituting a threat (i.e., impending harm), a loss (i.e., irreversible harm that has already been incurred), or a challenge (i.e., a stressor the individual is confident about mastering).

Figure 10.1 Coping Depicted as a Transactional Process of Appraising and Dealing with Demands. These appraisals trigger bouts of coping, defined as “cognitive and behavioral efforts to master, tolerate, or reduce external and internal demands and conflicts among them” (Folkman & Lazarus, 1980, p. 223), which utilize personal and social resources to solve the stressful problem or manage the individual's negative emotional reactions to it. These efforts produce coping outcomes, which, by feeding back to both the stressful event and individuals' reappraisal processes, can terminate or prolong the stressful transaction. According to this perspective, coping can be seen as a process that involves a wide variety of ways of reacting to and dealing with stressors that are organized sequentially, forming an interconnected action sequence or coping episode (Folkman & Lazarus, 1985).

Stress, Appraisals, and Coping Associated with Adjustment and Psychopathology A primary focus of research on stress and coping across the life span has been on correlating the many ways of coping with indicators of adjustment, such as internalizing behavior (e.g., depressive and anxiety symptoms or disorders) or externalizing behavior (e.g., behavior problems and aggression or conduct disorder). In fact, in 2000, Coyne and Racioppo reported that the cumulative number of publications with coping as a keyword totaled almost 25,000 articles, with the great majority focused on identifying stress and coping as risks or resources for mental health and adjustment. Literally thousands of additional studies have been published in the past 15 years. The size of the literature makes it difficult to achieve a comprehensive review, but a few trends are apparent. For example, in the most recent decades, contemporary research has focused on stress and coping as correlates of physiological functioning and physical health (Appleton, Buka, Loucks, Gilman, & Kubzansky, 2013; O'Leary, 1990; Penley, Tomaka, & Wiebe, 2002; Walker, Smith, Garber, & Claar, 2007), and even more recently, a small but growing number of studies focus explicitly on stress, coping, and psychopathology (e.g., Boxer, Sloan-Power, Mercado, & Schappell, 2012; Tolan, Gorman-Smith, Henry, Chung, & Hunt, 2002). Multiple reviews summarize much of this research (Bridges, 2003; Clarke, 2006; Compas, Connor-Smith, Saltzman, Thomsen, & Wadsworth, 2001; Compas, Orosan, &

Grant, 1993; Decker, 2006; Frydenberg, 1997; Nes & Segerstrom, 2006; Penley et al., 2002; Petticrew, Bell, & Hunter, 2002; Seiffge-Krenke, 2011; Taylor & Stanton, 2007; Wolchik & Sandler, 1997). Stressful Life Events Across these reviews, it is clear that major life stressors, including the death of a loved one, witnessing a traumatic event, or experiencing abuse by family members or others, are common experiences among children and adolescents, occurring for about 25%. An even greater number of children and adolescents experience repeated, sometimes daily, hassles related to school (e.g., fights or problems with teachers or academic performance) and interpersonal relationships (e.g., conflicts or problems with parents, siblings, and peers; Donaldson, Prinstein, Danovsky, & Spirito, 2000). Both significant life events and daily hassles have been associated with increasing symptoms of psychopathology over time, including depression, anxiety, and delinquent behavior (Compas, 1987a,b; Compas et al., 2001). Although these associations are usually small to moderate in most studies, associations are much stronger when specific stressors are examined, such as peer victimization (Harper, 2012; Kochenderfer-Ladd & Skinner, 2002; Zimmer-Gembeck, Hunter, & Pronk, 2007; ZimmerGembeck et al., 2013), friendship and romantic formation or dissolution (Nieder & SeiffgeKrenke, 2001; Seiffge-Krenke, 2011; Zimmer-Gembeck, Siebenbruner, & Collins, 2001); racial discrimination (Berkel et al., 2010; Brittian et al., 2013; Pascoe & Smart Richman, 2009; Umaña-Taylor & Updegraff, 2007), or community violence (Fowler, Tompsett, Braciszewski, Jacques-Tiura, & Baltes, 2009). Despite clear evidence that the experience of stressful life events is a risk factor for children and adolescents, it has been widely acknowledged that, in order to understand the development of psychopathology, it is important not only to attend to the intensity and chronicity of stressful events but also to take into account individuals' appraisals of stress, their coping responses, their feelings of efficacy in being able to carry out successful coping efforts, and their personal and social resources for coping (Moos & Holohan, 2003; Taylor & Stanton, 2007). In the remainder of this section we summarize and build on previous reviews linking stress and coping with psychopathology and positive adjustment, emphasizing studies of childhood and adolescence. We begin with what is known about general coping categories and their associations with psychopathology. We then consider how psychopathology and adjustment are related to specific coping strategies, to cognitive appraisals of stressful events, to coping selfefficacy, and, more recently, to profiles or combinations of coping strategies and resources. We end with a summary of transactional models of the connections among stress, coping, adjustment, and psychopathology as well as findings from new and innovative research studies, using intensive repeated measures, which provide evidence of reciprocal linkages between coping and psychopathology. Challenges to Summarizing Research on Coping In integrating studies of coping, we faced three challenges that have often been noted in the literature. The first was to figure out how to compare the many ways of coping examined

across different studies—which show little consistency and typically employ different or partially overlapping lists of ways of coping. Hundreds of ways of coping have been identified, which have been combined into dozens of instruments to assess them, primarily through standardized self-report questionnaires, written open-ended responses, or open-ended interviews, but also via others' reports of coping (e.g., spouses, parents, teachers, friends) and, less frequently, diary methods and observation (Skinner, Edge, Altman, & Sherwood, 2003). In efforts to manage coping's apparent complexity, many researchers who examine the links between coping and psychopathology have boiled these variegated response options down to two categories, sometimes referred to as problem-focused versus emotion-focused, engagement versus disengagement, approach versus avoidance, or active versus passive (and described in more detail in subsequent sections). A second challenge, given their potential overlap, was to distinguish between ways of coping and psychopathology, both conceptually and in terms of measurement. Maladaptive ways of coping sometimes are so closely connected to psychopathology that researchers have questioned whether coping assessments tap symptoms rather than predictors of disorder (Compas et al., 2001). Moreover, studies sometimes use measures that confound personal or social attributes, stress, and psychopathology, so that, for example, interrelations between coping and psychopathology might be explained by overlapping item content related to levels of stress and emotional distress (Austenfeld & Stanton, 2004; Coyne & Racioppo, 2000; Lazarus, 2000; Park, Armeli, & Tennen, 2004). As a result, in measurement work, researchers have had to carefully remove items from coping inventories that overlap with measures of emotional distress, and internalizing and externalizing behavior (Ayers, Sandler, West, & Roosa, 1996; Connor-Smith, Compas, Wadsworth, Thomsen, & Saltzman, 2000; Treynor, Gonzalez, & Nolen-Hoeksema, 2003). A third challenge to integrating work on coping stems from the broad array of stressor domains (e.g., medical conditions, victimization, environmental disasters) that are covered in current research on children and adolescents. It is now clear that conclusions about whether certain kinds of coping responses are adaptive versus maladaptive across a range of mental health outcomes (e.g., Kendall & Terry, 2008) depend heavily on the type of stressor with which children and adolescents are dealing, particularly on identifiable features of stressful events, such as their severity or perceived controllability (Skinner & Zimmer-Gembeck, 2011; Zimmer-Gembeck & Skinner, 2011). In fact, in recent years, this conclusion seems to be implicitly acknowledged in the coping literature as research has become even more differentiated and focused—as reflected in the increasing proliferation of special population studies in which researchers often focus on a single specific stressful health condition, chronic situation, or acute event, with the aim of explaining one or many potential adjustment outcomes.

Links between Broad Categories of Coping and Psychopathology Problem-Focused and Emotion-Focused, and Approach and Avoidance Coping One of the most commonly known broad categorizations of coping is the differentiation of strategies that are primarily problem-focused from those that are more emotion-focused

(Lazarus & Folkman, 1984). Problem-focused coping is usually defined to include strategies enacted in an attempt to modify or directly confront the stressful event, such as problem solving and direct action. Emotion-focused coping is usually defined to include responses that serve the purpose of managing emotional reactions to stress, such as social withdrawal, distraction, and emotional venting. Studies that examine the association of problem-focused and emotionfocused coping categories with psychopathology have considered a wide range of outcomes, most frequently depression, anxiety, loneliness, suicidal ideation, self-esteem, and positive well-being. A smaller number of studies have assessed additional outcomes, such as stress reduction, physiological reactions, or physical health (e.g., Appleton et al., 2013). In reviewing the literature on problem-focused and emotion-focused coping and adjustment among children and adolescents, Losoya, Eisenberg, and Fabes (1998) noted two general trends: (1) problem-focused coping strategies are associated with fewer emotional and behavioral problems and greater social competence, whereas emotion-focused coping is generally associated with more internalizing and externalizing symptoms; and (2) these findings are not uniform across studies and generally depend on the type of stressor or features of the stressor. Subsequent research confirms both of these trends. For example, in one recent study, the use of more problem-focused coping responses was correlated with fewer symptoms of mental health disorders (Li, DiGiuseppe, & Froh, 2006), but other studies found no such associations (e.g., Horwitz, Hill, & King, 2011). Findings are slightly more consistent across studies of emotion-focused coping, in that greater use of these strategies is correlated with elevated symptoms of mental health disorders (Horwitz et al., 2011; Rafnsson, Johnson, & Windle, 2006). However, a recent study by Krattenmacher et al. (2013), which also found that problemfocused coping was associated with better mental health, draws attention to the different forms that emotion-focused coping can take, producing a more differentiated pattern of associations with psychopathology (see Stanton, Danoff-Burg, Cameron, & Ellis, 1994, for similar conclusions in work with adults). In a study of 214 adolescents coping with their parents' cancer, some emotion-focused strategies, specifically those that also had an approach or active component, such as cognitive reappraisal, were associated with better mental health, whereas those strategies that had the function of venting emotions or of avoiding emotions or stressors were associated with worse mental health (see Austenfeld & Stanton, 2004, for more detail on emotion-focused coping subtypes). In summary, the general evidence points to the positive role of problem-focused coping and the negative role of emotion-focused coping in adjustment. Yet such findings are often inconsistent across studies, potentially because the broad categorization of problem-focused versus emotion-focused may not always adequately discriminate the pattern of coping responses that account for increasing mental health problems or effective resolution and adaptation. Although the broad categorization of problem-focused versus emotion-focused coping is probably better known (Lazarus & Folkman, 1984), the most commonly used categorization of coping in recent studies of children and adolescents refers to general styles of approach (sometimes also called active or engagement coping) and avoidance (sometimes also referred to as disengagement coping; e.g., Causey & Dubow, 1992; Compas et al., 2001; Ebata & Moos,

1991; Jaser et al., 2007; Lengua & Stormshak, 2000). Definitions of approach coping, which often share many of the same responses as problem-focused coping, include cognitive or behavioral efforts to manage the stressor. These are usually measured as problem solving, cognitive reappraisal, information or support seeking, and taking concrete action. Avoidance coping includes both cognitive and behavioral responses that serve the function of avoiding the distressing event or circumstances. Most often this is measured as distraction from the stressor, ignoring the situation, denial or minimization, withdrawal, escape, and/or wishful thinking. When reviews and the latest research are considered, there is substantial evidence that, among children and adolescents, approach and active forms of coping are associated with positive adjustment as well as with fewer symptoms of mental health problems. Such findings are generally consistent with the findings for problem-focused coping strategies (Compas et al., 2001; Fields & Prinz, 1997; Holahan & Moos, 1991; Lengua & Stormshak, 2000). Evidence is especially strong when measures of approach coping tap the use of these responses relative to avoidant forms of coping. In contrast, avoidant forms of coping are very frequently associated with risks for heightened negative outcomes, such as greater psychological distress and elevated behavior and other externalizing problems (Krattenmacher et al., 2013; Rohde, Lewinsohn, Tilson, & Seeley, 1990; Seiffge-Krenke & Klessinger, 2000). Three reviews have summarized evidence about the associations of approach and avoidant coping categories with psychological adjustment among children and adolescents. In the first review of 21 such studies (Fields & Prinz, 1997), approach/active strategies, such as seeking social support and direct problem solving, were associated with greater social competence and fewer internalizing and externalizing behavior problems. In a second review, by Compas et al. (2001), two classes of coping—“active” coping (which typically entails problem solving, problem-focused support, and/or information seeking) and “engagement” coping (which includes not only problem solving but also constructive emotional expression and support seeking) were associated with fewer internalizing problems in 29 of 40 studies, fewer externalizing problems in 15 of 18 studies, and greater social and academic competence in 17 of 20 studies. Within the broader factors of problem-focused and engagement coping, the individual ways of coping that have been found to be linked most closely to better functioning include problem solving and positive reappraisal. The Compas et al. (2001) review also noted that two kinds of coping, “emotion-focused coping” (which usually includes expression of negative affect, denial, and/or wishful thinking) and “disengagement coping” (which typically entails problem avoidance, cognitive avoidance, and social withdrawal) seem to be detrimental to positive functioning, in that they are linked to elevated internalizing and externalizing problems. The individual ways of coping that have been found to be most consistently associated with poor adjustment include cognitive and behavioral avoidance, social withdrawal, wishful thinking, self-blame, resigned acceptance, emotional discharge, venting, and self-criticism (Compas et al., 2001). Active Coping In the third review, Clarke (2006) conducted a meta-analysis to summarize effect sizes from 40

studies of youth age 21 and under focusing on the connections between active/approach coping in response to interpersonal stressors and adjustment. In analyses of externalizing behavior, internalizing behavior, social competence, and academic performance, only the association between active coping and social competence was significant, with a mean effect of .11 (confidence interval .01–.21). The effect of active coping on internalizing was moderated by age, however, with a stronger, and surprisingly positive, effect found in adolescents (.14, 17 studies) compared to preadolescents (–.04, 10 studies). Thus, it appears that, on average, active coping with interpersonal stress may have a small link to more internalizing symptoms among adolescents, but not among preadolescents. It may be helpful, however, to interpret this finding of age moderation in connection with the controllability of the stressor. Clarke found that the type of stressor was a particularly important moderator of the effects of coping: Active coping in response to controllable events was positively associated with healthy functioning (.05 to .22) but negatively associated when used in response to uncontrollable events (–.06 to –.19). In particular, utilization of active coping is especially adaptive in situations that children and adolescents expect to be amenable to personal control, such as school-related tasks. When situations are objectively less controllable, such as the experience of parental conflict or medical procedures, active coping may not be associated with adaptation or positive outcomes and sometimes may actually predict mental health problems. In these situations, some kinds of coping that reduce the experience of stress, such as self-distraction, may be more adaptive. Primary and Secondary Control Coping Following up on studies suggesting that the effects of different ways of coping depend on whether target events are amenable to personal control, recent research has focused on the (objective, researcher-defined) controllability of environmental stressors (Forsythe & Compas, 1987). Most coping researchers now agree that the use of active coping is associated with fewer externalizing problems and greater social competence, but only when it is deployed in response to controllable stressors as compared to uncontrollable ones. This finding has been extended to understanding other broad categories of coping and psychopathology, specifically categories like primary and secondary control coping (Rothbaum, Weisz, & Snyder, 1982), in which the term primary control coping refers to attempts to change the stressful situation through typical problem-focused, active, and approach ways of coping, whereas secondary control coping refers to strategies that allow people to accommodate to events, placing less effort on trying to change them. Secondary control strategies include cognitive reappraisal, focus on the positive, distraction, and willing acceptance (Connor-Smith et al., 2000). In general, research suggests that it is more adaptive to use primary control coping when dealing with controllable stressors and secondary control coping when dealing with uncontrollable ones. For example, one study found that adolescents were lower in internalizing symptoms when they used more secondary control efforts (i.e., accommodation) to adapt to the stress of having a depressed parent (an uncontrollable event; Langrock, Compas, Keller, Merchant, & Copeland, 2000). Extending this finding, another study reported that adolescents with depressed parents were better adjusted when they used active secondary control, such as positive thinking and cognitive restructuring, in response to situations that were less

controllable (e.g., family stress) and used active primary control, such as problem solving and emotional modulation, in response to more controllable situations (such as peer stress; Jaser et al., 2007). Finally, in a third study of 204 children and adolescents and their experiences of uncontrollable stressors (parental depression and interparental conflict), those who used more secondary control coping strategies of acceptance, distraction, cognitive restructuring, and positive thinking were lower in both child and parent reports of anxiety and depressive symptoms as well as lower in aggressive behavior (Fear et al., 2009). In summary, these reviews and recent empirical evidence illustrate that the fit between the situation and the use of approach or accommodative coping responses is an important indicator of likely adaptive or maladaptive outcomes among children and adolescents (and adults; also see S. M. Miller, 1992). Although it is still rare for studies to examine both the objectively defined controllability of stressful events and participants' own perceptions of controllability, these findings highlight the crucial importance of the capacity to accurately diagnose the controllability of stressors in order to implement approach or active coping or to draw on accommodative and distraction techniques depending on the needs of the stressful situation. Rather than relying only on approach and active coping attempts, children and youth, when faced with uncontrollable stressors, may benefit from accommodation strategies, such as positive reinterpretation and distraction, in order to deal with emotions, shield themselves from stress, promote acceptance, and allow them to move forward, all of which may be more healthy for concurrent and later mental health (Brandtstädter & Rothermund, 2002; Compas, Banez, Malcarne, & Worsham, 1991; Forsythe & Compas, 1987; Losoya et al., 1998; S. M. Miller & Green, 1985; Roth & Cohen, 1986). Connections between Specific Coping Strategies and Psychopathology Although less frequently a focus of research than studies examining broad coping categories, there have also been more fine-grained investigations of specific coping strategies and their links to psychopathology or competence (e.g., Decker, 2006; Khurana & Romer, 2012; Zimmer-Gembeck, Skinner, Morris, & Thomas, 2013). In general, results of studies of specific coping responses and adjustment or psychopathology are consistent with the findings for approach/active/primary, secondary control/accommodative, and avoidance coping and psychopathology. In particular, problem solving, a form of active and approach coping, is one of the most adaptive specific responses for maintaining good mental health in the face of stress. In a review of 12 studies of stress, coping, and personal attributes in adolescents with cancer, two specific coping strategies were consistently related to lower levels of depressive symptoms: problem solving and seeking alternative rewards (Decker, 2006). In other research with adolescents or young adults, help seeking (Gould et al., 2004), planning (Aldridge & Roesch, 2008), and positive reinterpretation (Stewart et al., 1997), as well as problem solving (Khurana & Romer, 2012), were each associated with greater competence or fewer symptoms of mental health problems. Yet there are also specific strategies that seem to prompt heightened or increasing adjustment problems. In particular, aggressive and ruminative coping are among the strategies that seem most detrimental to child and adolescent mental health. In one study, both were associated with

more internalizing symptoms and more peer reports of withdrawal or isolation from others (Sandstrom, 2004). In a second study, however, an even wider range of maladaptive coping strategies was identified (Horwitz et al., 2011). In this study of 140 adolescents seeking pediatric emergency services, five ways of coping and appraisals were measured, including denial, substance use, use of emotional support, behavioral disengagement, and self-blame. All were associated with more heightened depressive symptoms and more frequent suicidal ideation for boys, for girls, or for both. Unique coping and appraisal correlates of elevated depressive symptoms were behavioral disengagement, self-blame, and seeking emotional support. The one unique correlate of suicidal ideation was the greater use of emotional support, suggesting that emotional support seeking may also reflect level of distress. Other studies have found that behavioral disengagement, often assessed as behaviors that imply giving up or helplessness, is a risk factor for elevated depressive symptoms (e.g., Kaminsky, Robertson, & Dewey, 2006; Nolen-Hoeksema, Girgus, & Seligman, 1992; Wadsworth & Compas, 2002), and the use of distraction as a response to the stressor of discrimination is a risk factor associated with increased internalizing disorders over time (Brittian et al., 2013). Taken together, denial, rumination, aggression, or opposition in response to stress, as well as helplessness or disengagement, blaming the self, and seeking emotional support, are most indicative of greater symptoms of mental health problems among children and adolescents. Moreover, distraction may also be a risk for elevated symptoms, especially if it is used when stressful events are within personal control and active coping might be more productive. Although most studies have focused on older children or adolescents, there is also evidence that there are specific coping strategies associated with problems or competence in younger children. For example, Eisenberg et al. (1997), in their study of children 2 to 4 years of age who were reassessed again at ages 6 to 8 and 8 to 10, found that the use of destructive coping (more venting and avoidance/distraction, and less cognitive restructuring), especially when reported by teachers rather than parents, was negatively associated with concurrent and future levels of social competence and positively associated with externalizing symptoms and problem behaviors. Simultaneously, constructive coping (the use of instrumental coping and support seeking) was associated with greater social competence. In a study of 153 young children (K. A. Blair, Denham, Kochanoff, & Whipple, 2004), preschoolers higher in parentreported and teacher-reported passive coping (e.g., avoidance and denial of the problem) showed more signs of emotional and social maladaptation, and the use of passive coping sometimes exacerbated the effects of irritable or sad-fearful temperaments on externalizing or internalizing symptoms. In contrast, actively facing the problem (e.g., problem solving) was associated with fewer internalizing symptoms. Thus, even in very young children, there is emerging evidence that active coping rather than avoidant coping or passivity in response to stressful demands from the environment may be important for maintaining mental health and building social competence. Moreover, the capacity for adaptive coping may counteract temperamental challenges in the early years of life.

Do Subjective Appraisals of Stressful Encounters also Play a Role in Psychopathology?

According to the transactional perspective, individuals' subjective appraisals of stressful encounters are a key part of the stress and coping process (Lazarus & Folkman, 1984). Appraisals, including perceptions of threat, harm, loss, or controllability, as well as attributions of causality, responsibility, or blame, are important because they are imbued with emotion, prompt other cognitions, and guide subsequent behavior (Lazarus, 1994). In particular, appraisals can be motivating or demoralizing, can calm emotion or amplify distress, and can foster continued engagement with the environment or lead to disengagement and helplessness (Roesch & Weiner, 2001; Weiner, 1985). One complication within the research on appraisals, coping, and psychopathology, however, is that empirical distinctions between appraisals and coping are not always clear (Folkman, 1984). In some studies, causal attributions have been included as a way of coping (see Folkman, 1984, for a review). For example, self-blame, a form of internal attribution, has been examined as a coping strategy (e.g., Horwitz et al., 2011); and some multidimensional measures consider blaming others, sometimes referred to as projection, as a maladaptive way of coping (Skinner, Pitzer, & Steele, 2013). As Aldwin (2007) concluded, this research remains “muddled” (p. 182), and theories are needed that either clearly differentiate between appraisals and coping or explicitly posit that appraisals and reappraisals are integral parts of coping processes themselves. Regardless of the challenges facing researchers, it is now widely accepted that, along with children's coping responses, it is also important to consider their perceptions of control and other appraisals of stress in the development of psychopathology (Chorpita & Barlow, 1998; Skinner, 1995; Weems & Silverman, 2006). In general, the impact of stressful events appears to depend not only on the objective stressors themselves but also on subjective appraisals of stress, and these may vary from child to child. As would be expected, appraisals are important correlates of emotional reactions to stress and to mental health (Compas et al. 2001; ZimmerGembeck, Lees, Bradley, & Skinner, 2009). For example, the appraisal of a stressful event as more threatening has been associated with children's self-reported symptoms of anxiety, depression, and conduct-related problems following parental divorce (Sandler, Kim-Bae, & MacKinnon, 2000). One study illustrates the importance of appraisals in understanding how stress and coping are linked to psychopathology among children and adolescents. This study also raises the possibility that appraisals may be as important (or perhaps even more important) than coping for understanding the development of psychopathology. Lengua and Long (2002) examined threat and challenge appraisals, active and avoidant coping, and adjustment problems (depression, mother-reported internalizing and externalizing symptoms, and self-reported internalizing and externalizing symptoms) among 101 children ages 8 to 12 years old. Challenge appraisals were expected to be associated with active coping, as has been found in other research (Santiago-Rivera, Bernstein, & Gard, 1995). In contrast, threat appraisals were expected to be associated with maladaptive outcomes, because previous research had demonstrated that they were associated with depression, anxiety, and conduct problems in children experiencing parental divorce (Sandler, Tein, Mehta, Wolchik, & Ayers, 2000) and were found to be the conduit linking the stressor of interparental conflict to internalizing

symptoms (Grych, Fincham, Jouriles, & McDonald, 2000). Although higher stress levels, appraisals of greater threat and use of more avoidant coping were associated with elevated internalizing and externalizing symptoms in simple correlations, coping had no direct effects on internalizing and externalizing symptoms once the significant impacts on symptoms of stress levels, threat appraisals, and temperament were accounted for in a multivariate model. Other studies have found that the fit between appraisals and coping is important to consider when predicting psychopathology. For example, in a study of 76 adolescents who reported their coping and appraisals of the controllability of their cancer-related stress, the match between coping style and appraisals of controllability was associated with fewer depressive symptoms, less anger, less distress, and less anxiety (Sorgen & Manne, 2002). Thus, using more problem-focused coping when controllability was perceived to be high or using more emotion-focused coping when controllability was perceived to be low was associated with less distress and fewer symptoms. These findings show that, just as has been found for objectively defined controllability, subjective appraisals of controllability can be important to understanding the implications of coping for adjustment and psychopathology. However, to date, few studies have directly tested this goodness-of-fit hypothesis in children and adolescents (cf. Forsythe & Compas, 1987), and findings remain rather mixed even in studies of adults or university students (e.g., see Park et al., 2004, for a discussion and an example of mixed findings). Coping Self-Efficacy An additional set of appraisals that seem to be important to coping and mental health following encounters with stress can be found in global beliefs about one's capacity to cope effectively or successfully, sometimes referred to as coping self-efficacy (Smith, Calkins, & Keane, 2006; see also Galatzer-Levy, Burton, & Bonanno, 2012, for a similar construct referred to as coping flexibility). Building on the view that a sense of efficacy in coping promotes the subsequent use of constructive coping strategies, coping self-efficacy has been defined to include beliefs about one's own ability to manage stressful events themselves as well as perceptions of one's capacity to understand and adaptively regulate one's emotional reactions to stress. Adolescents who report more coping self-efficacy tend to cope more successfully with stressful events and to receive more interpersonal support (Sandler, Kim-Bae, et al., 2000; Sandler, Tein, et al., 2000). It is not yet clear whether coping self-efficacy should be differentiated from the stress and coping process, for example, by considering it to be akin to a personal or a social resource. However, because high coping efficacy also reflects a history of successful coping, it is probably best conceptualized as an indicator of self-perceived coping success, so that it captures not what young people actually do in response to stress but how effectively they feel they have done it. In such instances, coping efficacy might be considered a mediator that links coping to mental health outcomes, and such pathways have been documented. In one study of children (Sandler, Kim-Bae et al., 2000), approach coping was negatively associated with depression and anxiety, and avoidant coping was positively associated with emotional adjustment problems, with coping efficacy fully mediating these associations. The importance of considering both coping and coping efficacy when trying to understand

mental health was also suggested by a comprehensive study of children of alcoholic parents (Smith et al., 2006). In this study, in which all constructs were assessed via multiple reporters (mothers, father, teachers, children), children who had a history of more positive parenting (supportive parenting practices and consistent discipline) used more active and supportseeking coping and were higher in coping efficacy. Moreover, all of these factors were important in explaining why children differed in their levels of externalizing and/or internalizing symptoms. In particular, children's active coping and coping self-efficacy mediated associations between positive parenting and lower symptoms. Yet some findings depended on the reporter, on whether the outcome was internalizing or externalizing symptoms, and whether the child had an alcoholic parent. Causal Attributions Appraisals can also involve causal attributions or explanations for why an event occurred. These appraisals and attributions include not only estimations of the controllability of the event itself but also explanations about its cause, such as whether it was caused by the self or by something outside the self (sometimes called locus) or whether the cause was stable (versus unstable) or global (versus specific). Although no review of the role of attributions in coping during childhood and adolescence could be located, one meta-analysis was conducted that included 27 studies of adults who were coping with illness (Roesch & Weiner, 2001). When the results of these studies were quantitatively combined, internal attributions were found to be associated with more approach coping responses. Attributions that stressful events were more unstable and controllable were associated with more coping responses of all kinds, (approach and avoidance, problem and emotion focused). For example, patients who appraised their illnesses as stable and uncontrollable used more avoidance coping and were less well adjusted than those who viewed their illnesses as more unstable and controllable. Moreover, there was evidence that the effects of attributions on psychological adjustment were fully mediated via all coping responses (except for behavioral avoidance). Thus, in adults, at least when they are coping with an often uncontrollable stressor like illness, appraisals are important to understanding coping responses, adaptation, and psychopathology. The results of these studies raise the possibility that causality and attributional processes may also be critical for understanding coping prior to adulthood, or at least by late childhood or early adolescence. When focusing on these younger age groups, however, such findings should be considered only suggestive, given the cognitive changes that occur between childhood and adulthood. Perhaps the development of causal beliefs and attributions may explain some of the changes in coping that are found with increasing age (Band & Weisz, 1990; Zimmer-Gembeck & Skinner, 2011). It may also be the case that causal beliefs and attributions become more closely tied to coping processes with increasing age (Skinner & Zimmer-Gembeck, 2011).

Strategies for Emotion Regulation, Coping, and Psychopathology Coping is often aimed at regulating emotional experiences, either by changing one's own responses or by modifying the stressor that prompted the emotional reaction (Compas et al., 2014; Losoya et al., 1998). Overall, emotion dysregulation, usually assessed via measures

originally designed to assess coping, is thought to be a core feature of many forms of psychopathology (e.g., see Aldao & Nolen-Hoeksema, 2010, 2012; Aldao, Nolen-Hoeksema, & Schweizer, 2010; Webb, Miles, & Sheeran, 2012). Because of the clear conceptual and methodological overlaps, research on emotion regulation and the dysregulation of negative affect and their links to psychopathology often seem to be tantamount to studies of stress and coping, allowing some conclusions about coping and psychopathology to be drawn from such work. Not surprisingly, the basic findings from these two bodies of research are similar: Just as with studies of stress, coping, and psychopathology among children and adolescents (and even adults), complex interrelations have been found in the associations between psychopathology and the multitude of available emotion regulation strategies. At the same time, two general patterns can be discerned. First, three specific emotion regulation (or coping) strategies have been widely theorized to be protective against psychopathology: (1) cognitive reappraisal of the stressful event, defined as generating benign or positive interpretations or perspectives on stressful situations as a means of reducing distress (Aldao et al., 2010; Gross, 1998; John & Gross, 2004); (2) problem solving, which involves conscious attempts to change a stressful situation or contain its consequences; and (3) acceptance, which refers to willingly consenting to emotions as they are without judgment. Second, and also overlapping with research on coping, three emotion regulation strategies have been identified as risk factors for the development of psychopathology: (1) suppression, defined as the control and repression of unwanted thoughts and emotions and their expression; (2) avoidance, which involves evading an array of unwanted psychological experiences, including emotions, sensations, memories, and urges; and (3) rumination, which refers to repetitively focusing on negative and distressing experiences and emotions and their causes and consequences. Testing these expectations using a range of measures of coping and emotion regulation associated with stress, Aldao and Nolen-Hoeksema (2010) found that university students who relied more often on cognitive reappraisal and problem solving also reported fewer indicators of psychopathology (symptoms of anxiety, depression, and eating disorders), whereas symptoms were more elevated in students who relied more often on suppression and rumination. Moreover, associations were stronger between maladaptive strategies and symptoms than between adaptive strategies and symptoms. These patterns of findings were also confirmed in a meta-analysis of 114 studies that examined the associations between six forms of emotion regulation and four psychopathologies: Aldao et al. (2010) found that there were large effects of rumination on depression and anxiety; medium to large effects for avoidance, problem solving, and suppression; and small effects for reappraisal and acceptance. Three of these coping strategies—rumination, avoidance, and suppression—were associated with heightened levels of depression, anxiety, and (somewhat less strongly) substance use and eating disorders. The other three responses—problem solving, reappraisal, and acceptance—were associated with lower levels of symptoms. The strongest effect was for rumination, and the weakest effect was for acceptance. John and Gross (2004) also reviewed evidence of the links between adjustment and the emotional regulation strategies of reappraisal and suppression, concluding that research has predominantly supported the

positive implications of reappraisal and the negative implications of suppression. Most recently, many, but not all, of these associations were confirmed in a review of 306 experimental studies of emotion regulation and its effects on a range of outcomes including distress and adjustment (Webb et al., 2012). Much less research has examined the use of these emotion regulation strategies in children and adolescents. It may be that some of these strategies are not as widely used in younger age groups, given their cognitive complexity and the autonomous actions required to execute them. However, some research has been conducted, which has identified multiple emotion regulation deficits among children with internalizing and externalizing psychological difficulties (for a review see Zeman, Cassano, Perry-Parrish, & Stegall, 2006). Across these studies, deficits in emotion regulation and coping have been measured as biased or misguided appraisals of the threat and challenge of stressful events and excessive emotional reactions and displays. All of these deficits appear to be heightened among children with depression, anxiety, or bulimia nervosa (compared to children without these disorders), but patterns of deficits differ for children with internalizing versus externalizing disorders. In particular, children with elevated depressive and anxiety symptoms have been found to display poorer emotional awareness and emotional understanding, report lower self-efficacy regarding their ability to regulate their emotions and cope with stress, show a less well-developed emotional vocabulary, have difficulties expressing and regulating their anger, and more often display inappropriate expressions of sadness. In contrast, children identified as having elevated externalizing disorders seem to suffer from emotional undercontrol resulting in displays of high emotional intensity, evince poorer inhibition of anger as evidenced by facial and other displays, and respond to distressing events with less sadness than would be typical for other children.

Patterns or Profiles of Coping as Correlates of Psychopathology Emerging evidence suggests that coping strategies may be deployed in clusters or may interact in ways that heighten their effects on adjustment and psychopathology. Although recent studies have addressed this issue, the notion of coping patterns or clusters is not new. For example, Timko, Moos, and Mickselson (1993) argued for the importance of the relative or combined use of a range of coping strategies for psychopathology more than two decades ago. However, subsequent research has identified a number of important combinations of coping responses or coping resources. First, generalizing from findings across multiple studies, positive reinterpretation combined with acceptance may be especially beneficial for adjustment in uncontrollable situations, whereas positive reinterpretation combined with active coping may be especially beneficial to adjustment in controllable situations. Second, contrary to the idea that people rely on problemfocused versus emotion-focused coping strategies, multiple studies have found that coping strategies such as positive reappraisal, reinterpretation, or distraction (i.e., emotion-focused strategies) actually often occur in combination with or promote or antecede problem-focused coping, suggesting that calming emotion may promote constructive problem solving and that it may be this combination that aids good adjustment when facing stressful events, especially when events are objectively controllable (e.g., Mattlin, Wetherington, & Kessler, 1990;

Shimazu & Schaufeli, 2007; see also Folkman, 1984, for a review). Third, the negative effects of internalizing coping strategies, such as isolating oneself and emotional venting, may be weakened if they are used in conjunction with more active strategies, such as problem solving (Kingsbury, Coplan, & Rose-Krasnor, 2013). Fourth, social support may be most beneficial to relieving distress in situations of low personal control. In general, combinations or sequences of different coping strategies, theoretically or empirically derived, may turn out to be particularly helpful in understanding social and emotional adjustment following stressful events, both in the short and the longer term. At least two studies of adolescents have gone somewhat further in tackling this issue by implementing person-centered approaches (i.e., data clustering techniques) to better understand how the use of particular combinations of coping strategies may contribute to adjustment and mental health problems (Boxer et al., 2012; Tolan et al., 2002). In a first study, focused on youth living in impoverished areas of the inner city (Tolan et al., 2002), 372 adolescents 12 to 16 years old provided information twice over a 1-year period about their use of 12 ways of coping when facing difficulties or feeling tense. Stress (social) was also measured, and the psychopathological outcomes were youth reports of their internalizing and externalizing symptoms as well as teacher reports of the same symptoms. About 50% of the items formed seven coping categories (seeking support, venting emotion, avoidance by substance use, distraction, positive thinking, seeking guidance, and humor), and these were subjected to cluster analysis. Five clusters were accepted as the best representation of differentiation among adolescents in their person-level patterns of coping responses, and these were labeled: (1) support and guidance seekers, (2) minimal copers, (3) emotional substance users, (4) emotion-focused copers, and (5) complex copers. After controlling for demographic differences and stress levels between coping groups, the lowest concurrent internalizing symptom levels were found for support and guidance seekers and minimal copers. Over time, minimal copers showed greater increases in internalizing symptoms when compared to support and guidance seekers. In contrast, the highest concurrent externalizing symptom scores were reported by emotional substance users and emotion-focused copers; emotion-focused copers also had greater increases in externalizing symptoms over time when compared to support and guidance seekers and complex copers. Such findings suggest that reliance on seeking support and guidance, relative to other coping strategies, to manage stressful events during adolescence may serve to deflect the escalation of both internalizing and externalizing symptoms. In a second study, the focus was on coping with stress and violence among 131 adolescents living in a distressed metropolitan area (Boxer et al., 2012). Coping included externalizinginternalizing coping (which included such strategies as aggressive responses, emotional venting, and high emotional reactions) and distancing; and mental health was measured as symptoms of posttraumatic stress disorder (PTSD), externalizing, and internalizing. Three coping cluster groups were found: those identified as (1) low in all forms of coping, (2) high in internalizing-externalizing coping but low in distancing, and (3) high in distancing but low in internalizing-externalizing coping. The low coping group showed the highest levels of competence, whereas the high internalizing-externalizing group reported elevated symptoms of all disorders relative to the other two coping groups. At the same time, however, this study

illustrates the continuing problem of potential overlap in measures of stress, coping, and psychopathology. The measure of coping responses had obvious overlap with symptoms, and the levels of stress experienced by youth were assumed and not directly assessed, making it likely that the group of adolescents who reported the lowest levels of coping was also exposed to the fewest stressful events related to neighborhood violence. Once these methodological issues are addressed, however, person-centered analyses that identify coping patterns have much potential to inform our understanding of developmental pathways of psychopathology and resilience.

Transactional Models of the Links between Stress, Coping, and Psychopathology Partly because the direct associations between stress exposure (or adversity) and psychopathology have been rather small, transactional theories often posit complex roles for appraisals, coping, and coping resources in the processes that protect (or harm) children and adolescents at risk for psychopathology (Fields & Prinz, 1997; Lazarus & Folkman, 1994; Moos & Holahan, 2003; Skinner & Zimmer-Gembeck, 2007). At least four kinds of general models depicting the role of coping and its associated processes can be identified (see Figure 10.2): Coping has been conceptualized as a moderator, mediator, mechanism, and reciprocal process in the relations between stress and psychopathology. Because of such theoretical developments, studies of mediators and moderators have become the norm rather than the exception in coping research. As a result, there is research to support or refute all of these more complex views, as described in this section.

Figure 10.2 Four Models of the Role of Coping in the Processes that Connect Stress to Psychopathology, as a (1) Moderator; (2) Mediator; (3) Mechanism; and (4) Reciprocal Process. Coping as a Moderator, Mediator, and Mechanism One of the most straightforward models posits that all (or some forms) of coping are moderators that minimize, buffer, or exacerbate the negative effects of stress on adjustment or on the onset or relapse of psychopathology (Aldwin, 2007). Coping can be thought of as a stabilizing or destabilizing factor that helps maintain positive psychological adjustment during stressful periods or may explain why stressors lead to psychopathology. From this perspective, coping efforts should be most helpful when stress is severe and distress is high. Consistent with this notion, Holahan and Moos (1991) found that under high stress, personal and social resources (self-confidence, easygoing disposition, family support) were related to better psychological adjustment indirectly through their link to greater reliance on approach coping responses. Under low stress, only personal and social resources were related to adjustment (directly). When tested, moderation effects are often found, although not consistently across studies and not for all ways of coping. Nevertheless, some of the differences are consistent with theory and

conceptual frameworks. For example, in one study, moderation effects were found for problem-focused coping but only main effects of coping on psychological symptoms were found for emotion-focused coping strategies (Aldwin & Revenson, 1987). To explain this finding, the authors argued that emotion-focused coping may, to a large extent, reflect predispositional characteristics and that such characteristics may be more likely to show main, but not moderating, effects. In contrast, problem-focused strategies may be more situational specific—they are influenced by situational constraints and affordances—and, therefore, may have their effects by interfering or modifying the situation, which, in turn, has implications for adjustment. A second model (see Figure 10.2) views coping as a mediator or an adaptive process that itself is embedded in or shaped by stress and is the primary (or only) pathway through which stress exerts its proximal effects on adjustment and psychopathology (Aldwin, 2007). According to this model, one of the reasons that adversity has deleterious effects is that stressful life events trigger maladaptive coping in children and adolescents, which then puts them at risk for the development of psychopathology. For example, avoidant coping, often measured as denial or withdrawal, has been found to mediate between stressful circumstances and distress, on one hand, and concurrent or later adjustment, on the other. In one study, avoidant coping mediated the negative effect of living in a homeless shelter on women's depressive symptoms (Rayburn et al., 2005). Other studies have also found avoidant coping, particularly behavioral rather than cognitive avoidance, to mediate the impact of stress on adjustment (Barker, 2007; Holahan, Moos, Holahan, Brennan, & Schutte, 2005; Roesch & Weiner, 2001). In a recent study of two samples, one of early adolescents and the other of adults, followed for 7 or 12 months, respectively, rumination mediated the association of more stressful life events with increased anxiety symptoms in adolescents and adults and mediated the association of more stressful life events with increased anxiety and depressive symptoms among adults (Michl, McLaughlin, Sheperd, & Nolen-Hoeksema, 2013). However, other studies have found that other forms of coping do not mediate the effect of stress on adjustment or psychopathology (e.g., Aldwin & Revenson, 1987) and instead may have more direct effects on adjustment in some circumstances (such as the main effect of active coping with controllable events on more positive adjustment; Clarke, 2006). A third model (see Figure 10.2) posits coping as a mechanism through which protective factors exert their impact. Protective factors would include social resources, like social support, as well as personal resources, like optimism, personal control or mastery, self-esteem, or coping efficacy, which are hypothesized to proffer their protective effects at least in part by promoting constructive coping and discouraging reliance on maladaptive coping responses (Taylor & Stanton, 2007). From this perspective, in order to demonstrate the effects of coping per se, it would be essential to show that coping behaviors have unique effects on psychopathology over and above the effects of the preexisting personality and other characteristics that shape them (Carver & Connor-Smith, 2010). Complicating matters even further, there is new evidence that some coping responses interact with each other and that coping responses and resources may also interact with each other, suggesting that some personal traits or environmental conditions (i.e., coping resources) may

exacerbate or deflect the positive or negative effects of coping on outcomes (Dagan et al., 2011; Shimazu & Schaufeli, 2007; Taylor & Stanton, 2007). For example, Jacobsen et al. (2002) found that posttraumatic stress symptoms were most elevated among cancer patients who reported a combination of low social support and greater use of avoidant coping. Dagan et al. (2011) found that unsupportive spousal behavior was associated more strongly with distress among adult cancer patients when the patients were also low in personal control. Coping as Reciprocally Related to Psychopathology The fourth model (see Figure 10.2) views coping as part of a set of reciprocal processes binding it to psychopathology. According to these models, stress, coping, and psychopathology have bidirectional or reciprocal effects, whereby stress interferes with coping processes and contributes to maladjustment or psychopathological outcomes; at the same time, maladjustment and psychopathology generate later experiences of stress and undermine the development of coping responses and resources (Conway, Hammen, & Brennan, 2012; Hammen, 2005; Lazarus, 2000; Liu, 2013; Roesch & Weiner, 2001; Rudolph & Asher, 2000). Evidence from longitudinal studies has accumulated that documents both directions of these reciprocal effects. Results from time-lagged studies show that psychopathology can make it increasingly more difficult to cope adaptively with stress, just as multiple studies show that poor coping is associated with increasing adjustment problems and psychopathology. For example, multiple studies have shown that children (Zeman et al., 2006), adolescents (Littleton, Axsom, & Grills-Taquechel, 2011), and adults (Aldao et al., 2010; Roemer, Orsillo, & Salters-Pedneault, 2008) with elevated levels of depression, anxiety, or other forms of distress symptoms find it more difficult to use constructive coping strategies, such as cognitive reappraisal, acceptance, problem solving, and attentional redeployment. Depressed children also report using fewer strategies to regulate their negative emotions than nondepressed children (Garber, Braafladt, & Weiss, 1995), and adolescents with heightened depressive symptoms, relative to their nondepressed peers, report using more cognitive avoidance and resigned acceptance to cope with stressors, and less problem solving and positive reappraisal, all of which have been associated with the development of psychopathological symptoms (Ebata & Moos, 1991; Littleton et al., 2011). Such findings also extend to the flip side of these reciprocal relationships, as seen in the connections between coping and competence, whereby children rated higher in social competence by their parents also make more use of constructive coping strategies, such as problem solving and support seeking, compared to children rated as less competent (Zimmer-Gembeck, Lees, & Skinner, 2011). Although both directions of effects are not typically examined together in a single study, a few investigations have explicitly targeted and found bidirectional effects. In one such study, adults with more psychological symptoms (measured with the Langner 22-item Screening; Langner, 1962) experienced increasing stress and reported more maladaptive coping over time at the same time that maladaptive coping was found to predict increasing psychological symptoms over time (Aldwin & Revenson, 1987). In another study, reciprocal associations were found between distress related to a university mass shooting, coping responses, and PTSD symptoms, with maladaptive forms of coping (avoidance, social withdrawal, wishful thinking, and self-

criticism) related to increased distress over time, at the same time that levels of PTSD symptoms were associated with changes in coping over time (Littleton et al., 2011). As another example, adolescents' peer stress in the form of rejection by classmates has been associated with increasing depressive symptoms over time (Bagwell, Schmidt, Newcomb, & Bukowski, 2001; Garber et al., 1995; Panak & Garber, 1992; Parker, Rubin, Price, & DeRosier, 1995) and multiple, other forms of adolescent stress also have been associated within a chain of events that create increases in depressive symptoms over time (for a review, see Compas et al., 1993). However, it is important to point out that some longitudinal studies explicitly looking for bidirectional effects have found evidence for one pathway but not the reverse. For example, one study which showed that coping was associated with increasing or decreasing mental health symptoms over time did not find the converse temporal associations between earlier symptoms and changes in coping (Wadsworth & Berger, 2006); and another study revealed that some forms of psychopathology (e.g., PTSD) may prompt increasing maladaptive coping with stress over time, but such coping was not associated with escalation in symptoms over time (Littleton et al., 2011). In most cases, patterns of reciprocal relations between psychopathology and coping must be pieced together from multiple studies separately documenting complementary directions of effects. For example, studies have shown that adolescents' heightened depressive symptoms predict their reports (and their classmates' reports) of increasing peer stress over time (Zimmer-Gembeck et al., 2009) or other stress (Compas, Howell, Phares, Williams, & Giunta, 1989; Windle, 1992). Moreover, for adolescents, high levels of depressive symptoms can prompt increasing use of coping responses such as social isolation and avoidance (see Hammen, 1999, for a review), and, in turn, higher levels of withdrawal, isolation, or avoidance have been associated with increasing stress and distress among children (Gazelle, 2010; Prinstein & La Greca, 2002), adolescents (Littleton et al., 2011), and adults (Barker, 2007; Holahan et al., 2005; Shah, Gupchup, Berrego, Raisch, & Knapp, 2012). It is important to highlight the key role that avoidance coping seems to play in these bidirectional pathways between stress and depression. Daily Stress, Coping, and Psychopathology Stress, coping, and mental health also seem to be reciprocally linked across very short, even daily, temporal episodes. Because such series of interrelations can unfold so rapidly, stress and coping researchers have found the application of diary research designs particularly useful and informative. In these designs, participants often complete an initial assessment and then provide reports of their stress, coping, and other daily experiences for 5 or more days. Such studies tend to replicate the findings from cross-sectional and longitudinal research but also make clear that individuals differ in how rapidly they appraise stressful events, how often they use coping responses, and how necessary coping seem to be for maintaining positive emotion and well-being over time. For instance, in a study of (mostly female) university students' stress, coping, and appraisal of

stressor controllability, individual differences were found in the temporal patterns of controllability appraisals and coping (Park et al., 2004). Some individuals matched their appraisals to coping by using problem-solving coping strategies when events were appraised as controllable and using emotion-focused coping strategies when events were appraised as uncontrollable. However, others did not match appraisals and coping to the same degree. Although matching problem-focused coping to controllable situations was associated with better adjustment, matching emotion-focused strategies to stressors appraised as uncontrollable was associated with better adjustment only for a subset of participants. Moreover, the personality characteristic of neuroticism explained some of the individual differences in associations of controllability and coping with negative and positive mood across the 28 days of the diaries. In children, diary studies have also revealed the ways that coping and coping efficacy are associated with symptoms of mental and physical health problems over short periods of time. In particular, passive coping has been associated with more functional impairment, elevated depressive symptoms, and more somatization over a 5-day diary, whereas reports of coping efficacy have been associated with reduced symptoms of mental health problems, including fewer depressive symptoms and less somatization (Walker et al., 2007). Summary of Transactional Research Transactional research, which links reliance on different ways of coping (or emotion regulation) to different indicators of adjustment and psychopathology, is typical of the interindividual difference focus of most coping research with children and adolescents (as well as with adults). Also consistent across age groups is the pattern of findings suggesting that certain kinds of coping are positively linked to adaptive functioning and negatively to problems and disorders. These methods include active approach or engagement strategies like problem solving, positive reappraisal, constructive expression of emotions, and acceptance. In contrast, other ways of coping seem to show the reverse pattern of associations, that is, they are correlated positively to indicators of psychological problems and disorders and negatively to adjustment. These methods include avoidance and disengagement strategies, such as cognitive and behavioral escape, rumination, helplessness, suppression, social withdrawal, wishful thinking, emotional discharge, resigned acceptance, and self- and other-blame. Several kinds of appraisals and causal attributions also show consistent links with competent functioning and psychopathology through their effects on approach and avoidance coping. Appraisals of challenge, controllability, and attributions of stressful events to internal and controllable causes generally predict higher levels of approach coping and lower levels of helplessness and, through these, higher levels of competent functioning and lower levels of depression and externalizing behaviors. Appraisals of threat, uncontrollability, and attributions of negative events to internal, stable, and global causes generally predict higher levels of avoidance, self-blame, and ruminative coping, which in turn typically are connected with higher levels of disorder and lower levels of competence. Particularly important seem to be appraisals of coping efficacy, or individuals' confidence that they can deal successfully with the problems, setbacks, and negative emotions entailed in stressful experiences.

The study of subjective and objective estimates of the controllability of stressful situations has demonstrated that it is the match between situational demands and ways of coping that both characterizes competent functioning and predicts effective outcomes. As depicted in the Serenity Prayer, individuals need “the courage to change the things I can” because active approach “primary control” ways of coping are a good fit for stressors (or aspects of stressors) that are controllable (like schoolwork); as well as “the serenity to accept the things I cannot change” because accommodative, positive reappraisal, acceptance, and distraction coping are a better match for stressors (or aspects of stressors) that are not under an individual's personal control (like interparental conflict). Although not studied as often, this pattern of findings also implies that individuals must develop “the wisdom to know the difference” and suggests that children and adolescents would benefit from having parents who are aware of the fine distinctions and differences and can assist them in learning to make these distinctions. Particular combinations or sequences of coping have proven to be associated with competent functioning: for uncontrollable events, cognitive reinterpretation accompanied by acceptance; and for controllable events, cognitive reinterpretation along with active forms of coping. These combinations seem to blend ways of dealing with distressing emotion with ways of guiding productive behavior, perhaps even sequencing them so that distress does not interfere with constructive action. Other combinations suggest that it is the balance between adaptive and maladaptive forms of coping that is typical of more positive functioning under negative conditions: The deleterious effects of coping strategies such as isolation or emotional discharge can be buffered by the use of problem-focused or engagement coping; and high levels of social support may help compensate for low levels of personal control. Taken together, these findings suggest that it may be most useful to examine interconnected patterns of situational demands, appraisals, and profiles of coping in order to better understand social and emotional adjustment following stressful events in both the short and the long term. Multiple complex conceptual models have been proposed to account for the role of coping in the connections between exposure to stress, risk, or adversity, on one hand, and the development of psychopathology, competence, or resilience, on the other hand. Coping has been posited to act as a moderator, mediator, mechanism, and reciprocal partner in these connections. Evidence has been found for all these roles, but not consistently. Constructive coping, especially problem-focused coping, does seem to be especially crucial when stress is high (i.e., it seems to act as a moderator), but other forms of coping, such as those captured in typical emotion-focused amalgams, seem to be harmful under most circumstances. Avoidance coping, especially behavioral avoidance, seems to be a key pathway (i.e., mediator), in that high levels of stress and adversity elicit more avoidance and social withdrawal, which in turn is likely to create problems for psychological adjustment and functioning. However, some studies suggest that approach coping may both mediate the effects of stress and exert direct effects, irrespective of stress levels. Coping also seems to be a particularly important mechanism through which a variety of personal and social resources (such as mastery or sense of control, optimism, self-esteem, and social support), which have been found to act as protective factors under conditions of adversity, have been found to exert their beneficial

effects on competent functioning and to buffer children and adolescents from symptoms of psychopathology. Some of the strongest support for complex models comes from accumulating evidence that coping and psychopathology are reciprocally linked. Although not often investigated in the same studies, research increasingly documents the operation of each direction or pathway of influence. In terms of the effects of psychopathology on coping, studies show that emotion dysregulation and poor coping are hallmarks of many kinds of psychological disorders. In terms of the effects of coping on psychopathology, longitudinal, experimental, and intervention research show that when children and adolescents display these out-of-control emotions and maladaptive coping responses, they generate stress for themselves and their families and so initiate or exacerbate psychological problems. When taken together, most findings suggest that associations among stress, coping, and psychopathology are often reciprocal, unfolding in positive and negative developmental cascades across situations and over time.

Critique of Individual Differences Research on Coping and Psychopathology Over the last several decades, tens of thousands of studies have examined the connections between a variety of ways of coping and a variety of indicators of psychological adjustment and disorder. Studies in this cast have multiple problems, but they can be optimized through the improvement of measurement and design strategies. In terms of measurement, overlap among indicators of stress, coping, and maladjustment continues to be problematic. Maladaptive ways of coping seem so closely connected to expressions of distress and indicators of psychopathology that researchers continue to question whether coping assessments tap symptoms rather than predictors of disorder (Compas et al., 2001). As a result, in measurement work, researchers need to continue to carefully scrutinize and remove items from coping inventories that overlap with measures of stress, emotional distress, and internalizing and externalizing behaviors (Ayers et al., 1996; Connor-Smith et al., 2000; Treynor et al., 2003). A second set of problems is presented by typical study designs. The vast majority of individual differences research looks at links between coping and psychopathology at a single time point. As a result, as underscored by Compas et al. (2001) in their review of this work, it is not possible to determine a specific direction of effects from these associations: Although it is tempting to infer that the use of engagement or problem-focused coping leads to more successful adaptation to stress, this interpretation is tautological to a certain extent. That is, these findings may simply indicate that children and adolescents that are more socially competent, who are less anxious and depressed, and who exhibit fewer conduct and disruptive behavior problems are better able to generate solutions to problems and to maintain a positive outlook when faced with stress. (p. 118) This situation can be improved by the use of designs that include multiple times of measurement—either short term, such as daily diary time-series studies, or long term, such as longitudinal studies. These kinds of designs are essential to allow researchers to examine reciprocal effects.

The optimal study designs that allow causal inferences (in either direction) continue to be experimental or intervention studies, in which researchers foster more constructive coping and then examine the effects of these improved coping strategies on subsequent psychological symptoms. Researchers are more limited in their ability to experimentally study the reverse direction. That is, researchers are not ethically able to induce psychopathology, but they can induce dysphoria or negative mood and then examine these effects on how people cope with challenges and setbacks. Most persuasive are programs of research that utilize both experimental and naturalistic longitudinal designs to examine both directions of effects (e.g., Nolen-Hoeksema, Wisco, & Lyubomirsky, 2008). Despite the many contributions of transactional studies of individual differences in coping processes and psychopathology, even the best of this research suffers from an intractable problem: Any individual differences approach is inherently incomplete with respect to an understanding of psychopathology, which (at least from the perspective of the interdisciplinary field of developmental psychopathology, Cicchetti & Toth, 2009) is by nature a developmental process. This means that key questions can never be completely answered by looking at snapshots of the relations between coping and psychopathology at single time points or even at multiple points in time. Coping, like psychopathology, competence, and resilience, is inherently and reciprocally developmental: Developmental level decisively enables and constrains each of the processes depicted in transactional theories of coping: appraisals, personal and social resources, ways of coping, outcomes, and all the feedforward and feedback loops that connect them. At the same time, episodes of coping contribute to development: Through the repeated process of dealing with problems and difficulties, children and adolescents generate resources and liabilities for dealing with subsequent stressful encounters. To fully realize the value of transactional perspectives and to fully recognize the role of coping in the onset and progression of mental disorder, mental health, or the growth of resilience, an important next step is the creation and use of developmental conceptualizations of coping.

Normative Developmental Perspectives: Coping as a Set of Basic Adaptive Processes that Are Reorganized with Age Normative developmental frameworks are grounded in the proposition that coping is a fundamental adaptive process that has evolutionary value in allowing people to detect, manage, and learn from potentially dangerous encounters (White, 1974). From this perspective, it becomes clear that coping has its roots in many functional systems designed to deal with threats and dangers. Because all these subsystems are likely to be activated by stress, developmental conceptualizations posit that the term coping refers to how all of them are coordinated and sequenced during stressful encounters. Hence, in the field of coping during childhood and adolescence, coping has come to be defined as “regulation under stress” (Compas et al., 2014; Eisenberg et al., 1997; Rossman, 1992; Sandler, Wolchik, MacKinnon, Ayers, & Roosa, 1997; Skinner & Wellborn, 1994; Skinner & Zimmer-Gembeck, 2007, 2009).

When coping is seen as the coordination of adaptive processes designed to detect and respond to challenges and threats (see Figure 10.3) or as “reactivity and regulation under stress,” it becomes clear that the coping system needs to accomplish four tasks: (1) to detect and interpret information about internal and external demands (threat detection and appraisal); (2) to prepare a response based on internal and external guides and capacities (action readiness); and (3) to execute a response by coordinating action tendencies with internal and external demands and resources (action regulation). Moreover, in order to develop, the coping system also needs (4) to recover and learn from stressful encounters.

Figure 10.3 Depiction of the Coping System as a Set of Fundamental Adaptive Processes Used to Detect, Respond to, and Learn from Encounters with Potential Challenges, Threats, and Dangers. From this perspective, it becomes clear that coping represents a complex multilevel system (see Figure 10.4), which extends from (1) the neurophysiological level, including psychobiological subsystems involved in detection and reactions to stress and the regulation of stress reactivity, most centrally the hypothalamic-pituitary-adrenocortical axis (HPA), the sympathetic-adrenal medullary (SAM) axis, the amygdala, and the prefrontal cortex (PFC), especially the anterior cingulate cortex (ACC); (2) the psychological level, including processes involved in stress reactivity and regulation, especially the attentional, emotional, and motivational subsystems; (3) the level of action, including subsystems that jointly generate action tendencies and that integrate and regulate them, especially the behavioral, cognitive, and meta-cognitive subsystems; (4) the social level, including participation in coping by social

partners as well as interpersonal relationships (such as with caregivers, other family members, teachers, friends, and peers) that scaffold the development of coping's many subsystems; and (5) the societal level, including the stressors that impinge on and the resources that are available to children and adolescents themselves as well as the societal stressors and supports that influence their social partners.

Figure 10.4 Integrative Multilevel Conceptualization of Coping as a Set of Interrelated Processes that Function on the (1) Neurophysiological, (2) Psychological, (3) Action, (4) Interpersonal, and (5) Societal Levels. After considering these five levels, the reciprocal connections between coping and development become more apparent. On one hand, development shapes coping: (1) by exerting

extensive effects on the bottom-up processes that are coordinated and integrated during stressful transactions, including neurophysiological and psychological (i.e., attentional, emotional, motivational, behavioral, and cognitive) features of reactivity and action tendencies, as well as (2) exerting extensive effects on the top-down regulatory processes used to coordinate and integrate them. Higher-order developmental organizations contribute to coping by determining the nature of the tools children and adolescents have at their disposal to detect, respond to, and learn from threats at different ages as well as the sociocultural scaffolds that they can count on to protect them (or leave them vulnerable) while their own coping capacities are developing. On the other hand, coping also contributes to development: The actual moment-to-moment transactions between individuals and stressful events can be considered “proximal processes” that act as engines of development (Bronfenbrenner & Morris, 2006). These episodes, in which individuals are trying to optimize the fit between environmental and intrapsychic demands (stress) and their internal and external resources and responses (coping), can be seen as creating a “zone of proximal development,” in which strategies for dealing with stress and negotiating emotions can be discovered, practiced, and consolidated (or overwhelmed and discarded). Hence, cumulatively, such transactions can contribute to the development of capacities and resources for improved coping, regulation, and resilience (Carver, 1998; Carver & Scheier, 1998; Luthar, 2006; Tedeschi, Park, & Calhoun, 1998). A view of coping as a basic adaptive process suggests that, at the most global level, one way of outlining the age-graded progression of coping, as well as its qualitative shifts, is to consider several broad developmental phases that are characterized by different mechanisms of detection, appraisal, reactivity, regulation, and learning as well as different kinds of participation by social partners. As we outlined in our earlier work: Infancy would begin with stress reactions governed by reflexes, soon to be supplemented by coordinated action schema; during this period, caregivers would carry out coping actions based on the expressed intentions of their infants (interpersonal co-regulation). During toddlerhood and preschool age, coping would increasingly be carried out using direct actions, including those to enlist the participation of social partners; this would be the age at which voluntary coping actions would first appear (intrapersonal selfregulation). During middle childhood, coping through cognitive means would solidify, as described in work on distraction, delay, and problem-solving; children would be increasingly able to coordinate their coping efforts with those of others. By adolescence, coping through meta-cognitive means would be added, in which adolescents are capable of regulating their coping actions based on future concerns, including long-term goals and effects on others. (Skinner & Zimmer-Gembeck, 2007, p. 128) Age-graded shifts in the basic tasks of coping (i.e., detection and appraisal, action tendencies and regulation, and learning) are shaped, on one hand, by the development of neurophysiological subsystems, and, on the other hand, by changes in the demands and resources provided by social partners, and especially caregivers. In this section, we use research on the development of coping itself and of other regulatory subsystems (including

attentional, emotional, and behavioral regulation), to trace the course of age-related changes in how these subsystems are triggered and coordinated in the face of stress. Because social partners are so critical to the development of all these subsystems, we also detail the role of caregivers (and later, other social partners) in the emergence and consolidation of these regulatory resources and capacities. We focus especially on how the caregiver's role in coping changes over development, from one in which caregivers are doing most of the coping for newborns based on their infants' expressed preferences, to one of direct participation, then cooperation, and finally acting as a resource and backup system to the relatively independent coping of which adolescents are capable by the time they reach emerging adulthood. (For more details on the normative development of coping, please see Skinner & Zimmer-Gembeck, in press.)

Normative Development of Coping during Infancy: Implicit Coping Because systems to detect and deal with threats are essential to survival, newborns come equipped with preadapted responses to carry out these tasks. Sensory, perceptual, and attentional subsystems aid in threat detection; emotional and motivational subsystems aid in threat appraisal and action readiness; and the motor system carries out actions to express distress or respond to danger. The initial systems that coordinate newborns' threat detection and responses are based on reflexes and other automatic processes triggered by the neurophysiological stress reactivity subsystems, including the SAM, the HPA axis, and subcortical neurological subsystems, like the amygdala (Izard, Hembree, & Huebner, 1987; Ohman & Mineka, 2001). Attachment and “External Coping” At the level of action, these physiological subsystems initially trigger automatic motor behaviors, such as startling, huddling, crying, and diffuse whole-body reactions to physical discomfort, novelty, constraint, and other sources of psychological displeasure or distress. These expressions of distress are not initially intended as communications, but they can be read and interpreted by sensitive caregivers and so can initiate cycles of contingent and responsive caregiving. As described by decades of work on the attachment system (Ainsworth, 1979; Bowlby, 1969/1982; Carlson & Sroufe, 1995; Kobak, Cassidy, Lyons-Ruth, & Zir, 2006), these call-and-response cycles are part of species-general evolutionarily adaptive mechanisms, in which caregivers are predisposed to respond to a newborn's distress by engaging with and soothing the infant and by figuring out what is wrong and taking action to change the stressful situation, guided by the infant's needs and expressed preferences. Such soothing and comforting, along with actually meeting newborns' expressed needs, can be considered forms of “external coping” in which the caregiver, using information provided by the infant, appraises the stressor, analyzes the “problem,” and responds with emotion regulation or constructive coping actions. For the infant, such interactions cumulatively create the experience of a safe and trustworthy environment, which may be an important foundation for the infant's ability to modulate physiological stress responses (Fuertes, Dos Santos, Beeghly, & Tronick, 2006; Nachmias, Gunnar, Mangelsdorf, Parritz, & Buss, 1996). From the

first days of life, infants are also learning from stressful encounters, using mechanisms described by operant and associative conditioning (Harman, Rothbart, & Posner, 1997), in which habitual responses through their repeated use become embedded in implicit memory. These mechanisms soon modify infants' stress responses from those of reflexes to those of action schema or habits (Rothbart & Posner, 2006), perhaps regulated by subcortical structures, such as the amygdala, that rely on cumulative experience and learning about the nature of the environment and its relation to the actions of the individual (Lewis & Todd, 2007). Based on repeated experiences in a secure relationship with a sensitive caregiver, these cumulative patterns of learning, perhaps stored in the amygdala as “hot” information, may continuously downregulate the HPA axis and the SAM. This repeated experience with sensitive responding from others may be one of the mechanisms through which infants' stress reactivity subsystems become tuned to their environments (Spangler & Grossmann, 1993; Spangler, Schieche, Ilg, Maier, & Ackerman, 1994). When infants develop in the context of a secure attachment relationship, these neurophysiological systems (especially the HPA axis) go into a period of hyporesponsivity by about 3 months of age (Gunnar & Donzella, 2002; Gunnar & Quevedo, 2007), resulting in a calmer and less stress-reactive state for the infant after the first few months of life (Lewis & Todd, 2007). Proximity Seeking as an Omnibus Coping Strategy As described by attachment theory (Kobak et al., 2006; Sroufe, 1996; Sroufe & Waters, 1977), a secure relationship with the caregiver is also accompanied by the development of an omnibus coping strategy over the first few months of life, namely, proximity seeking. This response, to which human infants are biologically predisposed, relies on biobehavioral systems that are visible first in reflexes and crying and then, based at least partly on sensitive caregiving, in intentional communications and focused actions, such as looking, reaching, and distinctive vocal patterns and differentiated crying. By the end of the first 3 months of life, infants have developed, from a diffuse set of undirected expressions of distress and reflexive reactions, a differentiated set of appreciations and action tendencies that are integrated with caregivers' responses to infants' signals, including caregivers' strategies for repair and comfort (K. C. Barrett & Campos, 1991; Holodynski & Friedlmeier, 2006; Kopp, 1989). Following all of these accomplishments, coping shifts from what has largely been external coping to emotion regulation and problem solving that is coordinated between the active infant and the sensitive caregiver, and so can be more properly labeled interpersonal coping. Internal Working Models and Coping Appraisals Infants also construct generalized expectations, which have been referred to as “internal working models” of relationships, and they are considered to be the rudimentary beginnings of successively more complex and differentiated self-systems (Connell & Wellborn, 1990; Deci & Ryan, 1985; Laible & Thompson, 1998). Such self-systems include generalized expectancies of safety and security, or that loving care is available when one is distressed (Lewis, 1997), generalized expectancies for contingency and dependability and a sense of mastery or efficacy in the face of environmental challenges (J. S. Watson & Ramey, 1972), and generalized

expectancies that one's preferences will be attended to and respected (Deci & Ryan, 1985). These kinds of generalized expectancies likely continue to downregulate stress reactivity by broadcasting benign (implicit) appraisals of potentially stressful environmental encounters. Development of Regulation and Rudimentary Coping Infants and toddlers make momentous early advances in their capacity for regulation, which parallel the maturation of the PFC, which is involved in the processing, intensity, and regulation of emotion during the first years of life (Grimm et al., 2006). At this time, infants exhibit nascent cognitive control, and they can manage working memory and inhibitory control tasks (A. Diamond, Prevor, Callender, & Druin, 1997). They also develop rudimentary skills that they can use to regulate their emotional displays (N. A. Fox & Calkins, 2003; Rothbart, Derryberry, & Posner, 1994; Ruff & Rothbart, 1996) and are better able to direct their attention toward or away from environmental events in order to regulate emotion and action (Kopp, 2002). By about 18 months, toddlers can engage in effortful self-regulation primarily through selfdirected attention and voluntary control of action (Feldman, 2009; Harman, Rothbart, & Posner, 1997; Kochanska, Philibert, & Barry, 2009; Ruff & Rothbart, 1996). Such attentional orienting can serve as a way to guide thinking, feeling, and behaving, and the increasing efficiency of skills in each area builds the capacity of the entire system. Emotion regulation strategies are also improved by the end of the second year of life. All of these improvements parallel what is known about the maturation of the PFC in these early years (A. Diamond, 2002) and are also facilitated by changes in vagal tone (Feldman, 2009). Overall, by the second year of life, there are marked improvements in focused attention, response inhibition, effortful control, and emotion regulation, all of which aid in generating adaptive responses to the novel and discomforting events often encountered by infants. The capacities to inhibit and to attend are adaptive for many reasons, providing a foundation for the capacity to respond to others and to comply with parental requests or other environmental demands. Goal-Directed Action and Early “Problem-Focused Coping” In challenging but not overwhelming interactions with social and physical contexts, infants also develop the capacity for problem-focused coping as they flexibly deploy and focus attention on their goal-directed actions (Braungart-Rieker & Stifter, 1996; Bridges & Grolnick, 1995; McCarty, Clifton, & Collard, 1999). Especially interesting are interactions in which infants cannot fully realize their intentions; these can trigger rudimentary “coping” actions—for example, infants may increase their exertions toward the goal, become more energized, and later may try out different action strategies (DeLoache, Sugarman, & Brown, 1985; McCarty et al., 1999). If goals are blocked, infants may direct energy toward removing the obstacle, or they may withdraw their efforts and switch to another goal. These coping interactions are developmentally useful: The goals created by the stress of not being able to immediately realize their intentions spontaneously coordinates infants' biobehavioral systems, both within the associated neurological subsystems (Lewis & Todd, 2007) and between the neurophysiological subsystems and infants' action subsystems. Such

interactions not only exercise or consolidate existing connections but also stretch infants' actions into a zone of proximal development, where new strategies and combinations are generated. In fact, a shifting role of social partners is to maintain a set of conditions for infants under which they can develop increasing regulatory resources and capacities. In the most general terms, doing this involves helping infants maintain a state of biological integrity and stability while nudging interactions toward the zone of “just manageable challenge,” that is, providing opportunities for exploration and focused interaction with intrinsically interesting objects and people, combined with the availability of supports on an as-needed basis. Social Referencing and the Emergence of Interpersonal Coping Infants also begin to tune their appraisals of novel or ambiguous events more to their caregivers' signals of distress or interest, in a process known as social referencing, which allows infants to consult their caregivers' “radar” over some distance in order to make decisions about the potential dangers present in new situations and the extent to which they should engage or withdraw from specific encounters (Fonagy, Gergely, & Target, 2007; Lewis & Ramsay, 1999). At the same time, infants become more capable of indirect coping (K. C. Barrett & Campos, 1991) in which they delegate coping actions to caregivers through the use of intentional communications designed to elicit desired outcomes, such as pointing at a desired object. Caregivers' sensitive responsiveness can be considered a kind of coregulation in which babies and caregivers are in good communication about how to deal with challenging, and potentially threatening, encounters (L. M. Diamond & Aspinwall, 2003; Hornik, Risenhoover, & Gunnar, 1987; Lewis & Ramsay, 1999; Sorce, Emde, Campos, & Klinnert, 1985). These interactions form kinds of coping packages, which infants can initiate using increasingly differentiated intentionally communicated signals of their distress, internal states, preferences, and goals. Such episodes, repeated thousands of times, begin to synchronize infants' internal experiences of distress with their external expressions, acknowledged and mirrored by caregivers through attunement. These appreciations, or appraisals, are in turn coordinated with caregivers' external actions to relieve distress, and infants' subsequent internal experiences of relief from discomfort and satisfaction of goals. As these packages are differentiated, based on the underlying problem (cause) and its emotional markers, two new developments emerge: (1) infants construct sets of distress appraisals and expressions that are actual representations of their genuine underlying physiological states, emotions, and motives; and (2) they build up a repertoire of constructive interpersonal coping strategies that are targeted at the actual problem and are effective in dealing with the stressor and in bringing comfort, relief, and motive satisfaction (Calkins & Hill, 2007). Using their emerging representational capacities, infants can also begin to form subjective representations of the contingencies between these elements, allowing the packages to be stored in implicit memory for later use, when triggered by similar problems or emotional markers of distress.

Normative Development of Coping during Early Childhood: Voluntary Coping Early childhood (about age 1 to 4 years) is sometimes referred to as a period of emotional action regulation, because emotional systems seem to be coordinating toddlers' appreciations and action readiness in stressful situations (Holodynski & Friedlmeier, 2006; Kopp, 2009). Detection and responses to challenges and threats are largely carried out by the emotional and intrinsic motivational systems, which generate approach or engagement reactions to objects, people, or events that young children find attractive, and avoidance or withdrawal reactions to ones they find unattractive, frightening, or repulsive (K. C. Barrett & Campos, 1991). Based on a history of experiences with sensitive interpersonal coping and the construction of secure internal working models, toddlers become able to tolerate increasingly higher levels of stress, showing equanimity and patience, perhaps based on benign implicit appraisals of stressful situations and expectations that episodes will be resolved favorably. They are also able to clearly communicate their genuine emotions and desires and are generally ready and willing to cooperate with caregivers in regulating emotions and dealing effectively with action problems. Caregivers participate directly in appraisals and coping of toddlers at this age, as described by Gottman, Katz, and Hooven (1996, 1997) in their depiction of “emotion-coaching” parenting (see also Calkins & Hill, 2007; Keenan, 2000; Kliewer, Fearnow, & Miller, 1996; Power, 2004; Sroufe, 1996). Caring adults help children identify and discuss differentiated emotions, as well as their causes, and jointly examine strategies for tolerating or alleviating them (aka strategies for emotion regulation or emotion-focused coping) (Dunn, Bretherton, & Munn, 1987; Kopp, 1989; P. J. Miller & Sperry, 1987). Such emotion coaching allows children to integrate their genuine internal experiences of distress with a differentiated vocabulary to accurately recognize and represent a variety of emotions (Denham, 1998; Malatesta, Culver, Tesman, & Shephard, 1989; Saarni, 1997), thus affording children access to the full range of their emotional experience, which provides crucial information when they are appraising the meaning of a potentially stressful event and when they are coping. Representational Capacities and Coping, and the Development of Extrinsic Motivation The development of representational capacities, which were seen in early forms in generalized expectations and action schemes, produces a major shift in emotion regulation and coping with stress during early childhood (Denham, 1998; Derryberry & Tucker, 2006). These capacities make it possible for toddlers to maintain goals over longer periods of time, to “plan” successively more complex action strategies before carrying them out, and to communicate in more differentiated and accurate ways about their goals, desires, and emotions. The joint representation of internal states and preferences along with external affordances and conditions allows these elements to be more effectively coordinated and, with practice, to become linked; these joint representations then lead to more coherent and goal-directed interactions with social and physical partners, even under conditions of greater challenge and demand (Fonagy et al., 2007).

At the same time, the growth of representational capacities and working memory also shape the development of the emotional system, which comes to include the “other-conscious,” or socially communicated, emotions of pride, shame, and guilt. These new self-conscious emotions, coupled with close relationships and the desire to please attachment figures, ushers in the development of the “extrinsic” motivation system, which allows toddlers to comply to requests from caregivers to inhibit the expression of prepotent behaviors and emotions or to show behaviors or emotions that they do not spontaneously wish to perform (Kochanska, Coy, & Murray, 2001). The prepotent bottom-up emotional action tendencies generated locally by the intrinsic motivation system and marked by emotions can now begin to defer to top-down cognitively represented goals from the extrinsic motivational system, for which no action readiness is spontaneously available. Toddlers' capacities to comply seem to depend not only on the quality of the relationship with the caregiver but also on the strength and direction of the prepotent action tendencies that are generated (Kopp, 2009). Caregiving and the Development of Self-Regulation in Coping To support this transition, caregivers encourage young children's use of language to express their distress and requests (“use your words”), even under increasingly stressful conditions, and begin to regulate young children's action based not only on children's desires but also on cultural norms and moral principles for appropriate behavior (Power, 2004; Tolan & Grant, 2009). The shift from heteronomous regulation (or compliance, guided by caregivers) toward autonomous regulation (or self-regulation, guided by the young child's core self) not only allows children to begin to become the agent of their own coping but also requires children to more actively and intentionally coordinate their coping efforts with the needs and desires of social partners (Eisenberg, Fabes, & Murphy 1996; Eisenberg, Valiente, & Sulik, 2009). This shift has been studied most thoroughly in research on the development of compliance, which focuses on parental demands and norms (Kopp, 2009) and the development of conscience (Kochanska, Forman, Aksan, & Dunbar, 2005), which folds moral principles and priorities into coping. The kinds of scaffolding that seem to be most effective in promoting compliance, the internalization of conscience, and autonomous regulation more generally, include several elements that create “coping episodes” for young children (Kopp, 2009). First, parents provide consistent demands for appropriate behavior, focusing almost exclusively on insistence about only a very small number of “true moral rules,” such as honesty, treating the self and others with kindness and respect, and taking responsibility for one's mistakes and messes. Violations of these principles create “interpersonal problems” for young children with their parents, preschool teachers, or peers. In helping children solve these problems, one key is warm and caring structure provided by trusted adults, who offer alternative appropriate means, both verbal and nonverbal, for children to express their true feelings and desires. Especially important, in promoting both compliance and eventual internalization, are induction strategies that support autonomy by acknowledging children's genuine goals and feelings, combined with explanations of the relevance and importance of prosocial actions to children's own goals (Hoffman, 1994). Such inductions allow children to internalize alternative means of expressing

feelings and regulating actions under stressful conditions. Executive Functions, Problem Solving, and Coping The study of conscious control, or executive functioning, suggests that the emergence of such voluntary self-regulation is the product of neurological and cognitive developments that increasingly allow children to create and resolve conflicts in action regulation (A. Diamond, 2013; Zelazo, Muller, Frye, & Marcovitch, 2003). Previously, potential conflicts were resolved automatically by prioritizing prepotent habitual responses generated by the intrinsic motivation and emotion subsystems. With increases in working memory and attention and improved awareness of their own goals and intentions, children are able to represent conflicting sources of regulation—such as two rules for behavior, a bottom-up urge and a topdown goal, or a habit and a current alternative intention (Nigg, 2006; Pennington & Ozonoff, 1996). Executive functions allow young children to internally mediate these conflicts, for example, to begin to inhibit prepotent responses and to show alternative nondominant responses, or to shift the guides for their action regulation from one set of rules or tasks to another (Best & Miller, 2010; A. Diamond, 2013; Nigg, 2006). The internalization of prosocial strategies, concern for others, and moral rules for use in regulating action are combined with the continued development of the capacity to search for effective means to reach desired goals, or problem solving. Although it has its origins in contingency detection and tertiary circular reactions during infancy, intentional problem solving as a cognitive and social process comes into its own during early childhood (Keen, 2011). If handled sensitively, “stressful” transactions with uncooperative peers and materials (such as blocks, games, sports, or artwork) can become laboratories for developing problemsolving skills—allowing young children, with the help of adults and peers, to identify and generate ideas for new means or strategies, imagine their consequences, select from alternatives, try them out, and note their actual effectiveness (Berg & Strough, 2010). Individual Coping as a Supplement to Interpersonal Coping One of the most important reorganizations of the coping system takes place with the emergence of volitional action regulation during early childhood. By lifting regulation off emotions and integrating emotions with volition, coping becomes less reactive—that is, less a product of local conditions and implicit motives, and more flexible:—more open to top-down influences from both social partners and internal sources. The primary shift during this developmental period is from interpersonal to intrapersonal coping. The coping packages (with their appraisals differentiated by causes and emotions and their action repertoires for satisfying motives and soothing distress) that were coproduced by the child and the caregiver are now increasingly handed over to the child, and “independent” coping becomes the young child's developmental task, while scaffolding from adults (e.g., caregivers and preschool teachers) creates a zone of proximal development. For some theorists, who define coping as comprising only conscious and volitional efforts (e.g., Compas et al., 2001), the development of voluntary action regulation marks the beginning of coping proper. Such a shift requires children to reroute the interpersonal appeals that they previously directed

to caregivers (with their emotional expressions of problems and desires), so that they are directed intrapersonally for satisfaction (i.e., at their own newly emerging sense of self), and to use the information contained in emotions and language that was formerly used to guide the actions of caregivers toward meeting the child's needs, so that it is now employed to guide their own actions in meeting their own needs (Holodynski & Friedlmeier, 2007). Hence, the coping repertoire that was previously enacted between children and caregivers eventually comes to be reconstructed in the domain of children's own voluntary actions. It is important to make clear, however, that intrapersonal or individual coping does not replace interpersonal coping; it supplements it. Young children still have access to interpersonal strategies, which they can initiate through support, proximity, or help seeking, and they are likely to fall back on these strategies when the stressor is severe or the children are highly distressed, tired, or otherwise impaired (Zimmer-Gembeck & Skinner, 2011). At the same time, parents begin to dole out their direct participation, judging for themselves whether children are capable of coping on their own with the particular stressor in the current condition (Eisenberg et al., 2009). Caregivers may encourage a few bouts of independent coping and, if they see that children are intimidated or overwhelmed, may add resources or participate in a round of interpersonal coping to see whether they can tip the children back from a sense of threat and toward the experience of manageable challenge. Caregivers continue to shape emotion and coping by protecting young children from events that are potentially overwhelming, by coaching via direct instruction, and by more general discussions of problems and emotions (Fabes, Eisenberg, & Bernzweig, 1990; Morales & Bridges, 1996; Thompson, 1990; Valiente, Fabes, Eisenberg, & Spinrad, 2004). It seems likely that the development of all of the constructive ways of coping emerge from interpersonal scaffolding—not only of prosocial ways of coping (such as accommodation and negotiation) but also ones that are not so obviously social, like strategizing and self-soothing, which may emerge from joint problem solving and the coaching of emotion regulation.

Normative Development of Coping during Middle Childhood: Reflective Coping A major reorganization in the coping system takes place sometime during the 5-to-7 age shift (Sameroff & Haith, 1996), when children begin to internalize the mental means of regulation and slowly become able to regulate their actions using reflective consciousness. As noted by many coping theorists (e.g., Aldwin, 2007, Aldwin, Skinner, Zimmer-Gembeck, & Taylor, 2011; Compas et al, 2001; Murphy & Moriarity, 1976; Skinner & Edge, 1998a; Skinner & Zimmer-Gembeck, 2007, 2009), this cognitive revolution ushers in a widely expanded repertoire of coping because it allows children to reconstruct and deploy mental forms of all the ways of coping that they have previously used on the plane of direct action. For example, instrumental action is supplemented by cognitive problem solving, behavioral distraction by mental distraction, physical escape by mental withdrawal, physical soothing by cognitive emotion regulation, and so on.

Regulatory Development and Coping, and Construction from Intrapersonal Coping The capacity for voluntary coping that emerged in toddlerhood and improves up until school entry coincides with advances and increasing efficiencies in attentional skills and shifting of attention, the capacity to inhibit responses, effortful control, and emotion regulation (Fan, Fossella, Sommer, Wu, & Posner, 2003; Rueda & Rothbart, 2009). Children also exhibit great improvements in working memory by age 4, and this continues to improve up until age 7 (Luciana & Nelson, 1998). Children become able to delay gratification for longer and display decreases in impulsivity (L. B. Jones, Rothbart, & Posner, 2003; Prencipe & Zelazo, 2005). Children who are better at response inhibition by the age of 4 are better able to focus and shift their attention and are less impulsive and prone to frustration (Gerardi, Rothbart, Posner, & Kelper, 1996; Zelazo, Reznick, & Pinon, 1995). All of these capacities serve them well when coping with stress. Thus, children show major advances in coping and self-regulation between the ages of 3 and 6, and all of these advances seem to be occurring at the same time as maturation is progressing in the dorsal ACC network (Geidd et al., 2004; Gogtay et al., 2004). Moreover, the average volume of brain activation is reduced in adults compared to children (Casey, Jones, & Somerville, 2011), which suggests that brain areas involved in executive functions are becoming more focal and efficient, at the same time as children increasingly recruit more sophisticated prefrontal systems of the brain for self-regulation (Casey et al., 2011; Luna et al., 2001; Luna & Sweeney, 2004; Rubia et al., 2000). The brain seems to be moving toward specialization and reduction in the time needed to process information and respond. Some researchers refer to this as “fractionating” to reflect the differentiated and distributed functioning seen in the adult brain (Baddeley, 1998; Tsujimoto, 2008). Therefore, development may mean an increasing fractionation of brain function from early childhood to early adulthood in which multiple higher-level brain processes draw on common areas of functioning and covary with each other before the age of about 7, but these same processes become fractionated beginning at about age 7 or 8. This change is also apparent in neural organization, in that functions move from diffuse to more focal and fine-tuned for performance (Casey, Tottenham, Liston, & Durston, 2005; Durston et al., 2006). The reconstruction of the coping repertoire on the mental plane is accomplished through a long process of internalization and transformation of the previously enacted coping system. This process is made easier if earlier bottom-up action tendencies are coherent and genuinely informative about children's authentic feelings, motives, and neurophysiological stress reactions; and if developing top-down self-systems for regulating action tendencies are sturdy and constructive, that is, undergirded by high levels of trust (and so are socially cooperative), mastery (and so are optimistic), and self-determination (and so are agentic). Because coping, by definition, involves actions in potentially threatening situations of personal relevance, the system of direct coping actions is likely reconstructed on the representational plane as emotion-laden “hot” cognition. The most obvious features to be internalized are

language components: of stress reactions (e.g., distress labels), appraisals (e.g., emotion and problem labels), and responses (e.g., sequences of actions). Emotionally expressive signs and emotion experiences are also internalized, which then become capable of triggering emotional reactions in the absence of direct physiological stimulation (Holodynski & Friedlmeier, 2006). Also internalized are attention and motivation—in which children come to represent the priorities and goals that matter to them (Derryberry, Reed, & Pilkenton-Taylor, 2003; Rueda & Rothbart, 2009), which can then serve to direct focused attention as well as to energize engagement and withdrawal action tendencies (Block & Block, 1980; Metcalfe & Mischel, 1999). Advantages of Mental Means of Coping The internalization of the coping action system allows the child to reconstruct, in his or her imagination, sequences of potential coping responses and to play out not only their likely success in accomplishing the child's goals during stressful transactions but also their likely motivational, emotional, and social costs and benefits. The capacity to conduct such mental coping exercises confers many advantages: Children can conserve energetic resources that would be expended (and avoid the risks that would be incurred) if they were to try out these options on the plane of action. Mental constructions also allow for a qualitative shift in flexibility, as a wider variety of possibilities can be considered, including novel sequences that have not yet been directly enacted on the ground. From the perspective of action regulation, the biggest improvement during this developmental period is the capacity to coordinate all the features of coping on the same level, namely, the cognitive level (see Derryberry & Tucker, 2006, for a similar description of self-regulation and self-organization). Reconstructions of bottom-up physiological reactions, emotions, attentional priorities, and intrinsic motivations, along with top-down recommendations from extrinsic motivations, relationship considerations, cultural and social norms, and moral principles, can now all be brought together to the same table of hot cognitive representations to be coordinated through active and reflective mental “discussion,” making explicit the conflicts and trade-offs that were previously dealt with through competing action tendencies or regulation by others. These conflicts, which both create stress and are exacerbated by stress, can be consciously recognized and thoughtfully negotiated. “Mental” Participation of Social Partners Such internal negotiations are made easier by caregivers who are sympathetic to the dilemmas children face in stressful situations and are willing to continue coaching and cooperating with their attempts to work them out (Denham, 1998; Eisenberg et al., 1998; Gross & Thompson, 2007; Kliewer et al., 1996; Neitzel & Stright, 2003). Adults' direct participation in coping episodes slowly recedes across middle childhood, as their active participation is successively replaced with “mental participation” through cognitively mediated reflective means, such as discussions, reminders, encouragement, suggestions, and structures (e.g., routines, rules, and rituals). If parents are able to continue providing high levels of warmth and caring, structure, and

autonomy support as children practice dealing with conflicts, dilemmas, obstacles, and problems, out of these stressful situations can emerge the autonomous regulation of action, in which children internalize and integrate moral rules and sociocultural demands with their true selves (Mesquita & Albert, 2007; Ryan, 1992) and so become able to use these guides to organize their action even in the absence of external monitors. It is important to note that, even though children are potentially capable of autonomous self-regulation, it can still be difficult to enact under stressful conditions, for example, in conditions of “temptation,” when strong prepotent conflicting action tendencies are activated (Metcalfe & Mischel, 1999). Role of Coping Attempts and Failures The capacity to mentally coordinate and integrate information from all coping subsystems represents a major shift in the operation of the coping system itself. And it is through repeated attempts to accomplish this mental coordination, with the cooperation of social partners, that these skills emerge, are practiced and consolidated, and eventually can be reliably executed under conditions of increasing stress (i.e., under conditions of increasing internal distress and external pressure). Continuing to be extremely important during this phase are the reactions of adults (e.g., parents, teachers, coaches) to children's failures and breakdowns, which are key experiences in the development of resilient “mindsets” (Dweck, 2008). Children need practice with mistakes and failures in order to learn how to tolerate and benefit from the negative emotions mistakes generate; to view failures constructively; to engage fruitfully despite anxiety, frustration, or shame; to take responsibility for mistakes and repair them; and to figure out how to help negative emotions dissipate. Mental Means Supplement the Coping Repertoire and Sturdy Coping Systems These new mental means of emotion regulation and problem solving, although built on direct coping actions, do not replace them. Children still have access to previous behavioral forms of coping, which they are likely to fall back on when stress or distress is high and when mental means are too difficult to execute or are not as effective as action means. During middle childhood, the child's mental coping repertoire is not only expanded and differentiated but also coordinated and integrated with the previous repertoire of action, interpersonal, and automatic strategies. Hence, a major developmental task during middle childhood is to learn how to flexibly deploy this repertoire over the course of a coping episode—based on the severity of the stressor, the changing internal conditions of the child, and the actual effectiveness of the previous sequence of strategies in dealing with the stressor and its emotional concomitants (for reviews see Decker, 2006; Eisenberg et al., 1997; Fields & Prinz, 1997; Losoya et al., 1998; Zimmer-Gembeck & Skinner, 2011). Middle childhood, with its pragmatic cognition and well-developed self-systems, is a developmental phase during which coping is often described as “sturdy,” especially if coping systems are built on constructive action tendencies, guided by (1) a sense of control and competence that is calibrated to actual performance in different domains and that encourages challenge appraisals and mastery responses to problems and obstacles; (2) feelings of security and relatedness that are calibrated to the prosocial actions of the self and the actual

trustworthiness of different social partners—these internal working models encourage both self-reliance in action and emotion regulation along with appropriate support seeking as needed to restabilize independent coping; and (3) feelings of self-determination and autonomy that are calibrated to the true self and that encourage both accommodation to reasonable demands and self-assertive cooperative negotiation when the child's own needs are not met by current conditions. This sturdy coping system may be one reason why middle childhood is seen as a particularly resilient period—a time during which children can withstand and sometimes even flourish under conditions of adversity (Masten, 2001).

Normative Development of Coping during Adolescence: Proactive Coping By adolescence, much of the brain structure and circuitry that allow for accurate stress appraisals and sophisticated coping responses are in place, but there are developmental changes in both the bottom-up and top-down processes that shape stress and coping during this age period. The hormones of puberty seem to reopen the underlying stress reactivity systems governed by the SAM and HPA axis so that they may become increasingly susceptible to influence from external stressors. The period of hyporesponsivity of these systems, which started in infancy, comes to an end, and a period of more intense emotional and motivational activation of the bottom-up, amygdala-regulated reactivity systems commences (Dahl, 2004; Gunnar & Quevado, 2007; Lewis & Todd, 2007; Spear, 2003). Some theorists characterize this phase as one of imbalance between strong bottom-up emotionally charged impulses and the still developing PFC which cannot yet effectively regulate these urges (e.g., Casey et al., 2011; Steinberg et al., 2006), making emotion regulation and the focusing of cognition and attention when affect is high quite challenging for youth until later in adolescence or even into the mid-20s. This imbalance may also make adolescents vulnerable, especially under stressful conditions, to the short- and long-term consequences of highly attractive but highly risky behavior in a variety of domains (Fischhoff, 2005; Michel, Kropp, Eyre, & Halpern-Felsher, 2005; Reyna & Farley, 2006; Steinberg & Morris, 2001; Zimmer-Gembeck & Helfand, 2008; Zimmer-Gembeck, Siebenbruner, & Collins, 2004). In terms of coping, these challenges to emotion regulation in highly emotional environments during the earliest years of adolescence provide one explanation for trends suggesting that, after improvements in most ways of coping across early and middle childhood, during early adolescence, some youth seem to use more of a variety of maladaptive coping strategies (e.g., aggression, Pelligrini & Bartini, 2001) and become reluctant to seek help from adults (Newman, Murray, & Lussier, 2001). Regulatory Developments and Coping At the same time, increasingly stronger top-down regulatory capacities are slowly added, as can be seen in the growing efficiency of adolescents' executive functioning, involved in skills such as inhibitory control, problem solving, planning, logic, reasoning ability, and understanding consequences (Best & Miller, 2010; Reyna & Farley, 2006) along with other neurological developments (Spear, 2003). The PFC and the dorsal system of the brain,

including parts of the ACC, have a prolonged course of development continuing into adolescence, which is consistent with research showing improvements in executive control and inhibitory skill during this time (Dennis, 2010; Luna, Padmanabhan, & O'Hearn, 2010). Inhibitory control is relevant for stress and coping because it is critical for inhibiting both responses to threatening events and negative emotions. By about 13 to 17 years of age, many adolescents show adultlike performance on inhibitory control tasks, although their performance continues to require more effort than from adults (Flair et al., 2007; Luna et al., 2010). Similar to inhibitory skill, the development of emotional control seems to shows a prolonged pattern of development, but it has not yet been examined as thoroughly as cognitive control and executive functioning. Hence, it is not known, for example, whether it is the generation of emotion that changes with age or the capacity to control or manage it. In the same vein, working memory, a central component of executive function, operates quite well early in life but also improves substantially all throughout adolescence (Luna & Sweeney, 2004). Nevertheless, adolescents' cognitive capacities have not yet fully matured. Studies of cognitive control using functional magnetic resonance imaging (fMRI) have generally identified two critical differences between adolescents and adults: Adolescents have more difficulties with response inhibition, and their working memory performance may not yet be as fully developed (Luna et al., 2010). Because voluntary planned behavior depends on advanced cognitive skills, such as retaining goals online via working memory, preparing and planning responses, and suppressing task irrelevant responses (response inhibition), it seems clear that these types of tasks challenge adolescents' coping responses more than adults. Although development and participation of the PFC may be partial explanations for these improvements in cognitive control and the increasing efficiency of response inhibition and working memory, all of these advances in functioning also are supported by the increasing integration of diverse brain systems outside the PFC as children get older (Luna et al., 2010). As Luna et al. (2010, p. 109) described: Overall, these results imply that an important part of development is the process of specializing and segregating circuitries that support task ability, response state, and default processing. The ability to utilize cognitive control to perform a response, the ability to retain a response state, and to suppress internal thoughts improves with development, as the circuits supporting these distinct processes become independent. These suggest age-related improvements in white matter connectivity but also functional integration as seen in spontaneous waves of synchronized activity (Fair et al., 2008; Uhlhaas et al., 2009, p. 109) Metacognition and Coping These developing executive functions and the cognitive abilities of formal operations should result in a major reorganization of the coping system, brought about by the emergence during adolescence of metacognitive capacities, that is, the capacity to think about one's own thinking and cognitive activities (see Compas et al., 2001, for a review; also see Frydenberg, 1997; Kuhn & Franklin, 2006; Seiffge-Krenke, 1995). Metacognitive capacities supplement the

mental system of coping that was internalized during middle childhood, by integrating it with the adolescents' capacity to reflect on, evaluate, critique, and improve their own coping. Metacognitive capacities open the way for increasing and intentional self-regulation of all the coping families, such as problem solving (as seen, for example, in increasingly self-regulated learning), reflective emotion regulation, and the exercise of volition through identified selfregulation (Band & Weisz, 1990). All of these capacities combine to enable more proactive coping (Aspinwall & Taylor, 1997), in which regulation under stress comes to include not only a concern for current internal action tendencies and external opportunities but also future, longterm considerations. Coping responses that seem attractive in the short term come to be informed by graphic imaginations of the future, such as the possibility of being physically or psychologically hurt; of getting caught; of losing valued outcomes, relationships; and even one's own self-evaluations. As with mental means of regulation, these newly developing metacognitive capacities allow adolescents to differentiate and consider any aspect of the coping system, including appraisals, emotions, desires, action tendencies, and regulatory strategies. For example, metacognition allows adolescents to reflect on their interpretations of complex emotional states, their strategies for regulating them, and their potential to enact alternative strategies that might be more constructive. Moreover, “meta-emotions” allow adolescents to have emotional reactions to their emotional reactions, for example, feeling embarrassed about being afraid. Adolescents' increasing capacities for reflection, although generally positive, may also help to explain the increases seen during this age and in ways of coping focused on self-criticism and rumination and in reluctance to seek support (Zimmer-Gembeck & Skinner, 2011): The same capacities that allow adolescents to reflect on their emotions allow them to criticize their emotions; the same capacities that allow adolescents to plan for the future make them more vulnerable to worrying about the future; and the same capacities that allow adolescents to recognize they need help can also trigger concern that others may lose respect for them if they seek help. Identity Development and Coping Adolescent metacognitive capacities may also exert an influence on the nature of the self that is participating in regulation, as suggested by the fact that a major task of adolescence and young adulthood is identity development (Kroger, 2007; Meeus, 2011). The facets of the self that were differentiated by domain and calibrated to actual performance during middle childhood can now be thought of as parts of a whole that can be compared to each other and to the criteria of social norms and demands, with evaluative consequences for the self. To the extent that, in the process of identity construction, these disparate parts can be integrated successfully with each other, with the genuine self, and with the reflected regard of trusted others, a positive and coherent sense of self and identity can be achieved (Harter, 2012) that allows identified and autonomous self-regulation to become the dominant mode of coordinating bottom-up action tendencies and top-down considerations, even under increasingly higher levels of temptation and stress. Coping Flexibility, Attunement, and the Importance of Social Partners

These emerging metacognitive capacities may also help to explain the increases in coping flexibility that are seen during this developmental phase (Babb, Levine, & Arseneault, 2010), which is perhaps most noticeable in adolescent support seeking (Zimmer-Gembeck & Skinner, 2011). Adolescents become more differentiated and flexible about the kinds of support they request and the person to whom they turn for these social resources, seeking out parents, friends, teachers, or other adults or peers, and asking for comfort, distraction, advice, or instrumental help, depending on the local conditions: the domain and specific problem, the social partner's authority over the situation, and the kind of supports they need. With this new level of representation, it seems that the entire repertoire of previous forms of interpersonal coping can be integrated into the intrapersonal repertoire of support seeking, eventually enabling youth to initiate, access, and guide the deployment of available social resources while coordinating them with their own self-reliance and independent coping. This same pattern, in which coping strategies become more differentiated and more selectively applied and coordinated with local internal and external conditions, can also be seen in other families of coping—even in those more focused on independent or cognitive coping. For example, youth are increasingly likely to use problem solving in response to stressors which they perceive as controllable (e.g., academic problems) whereas they are more likely to use accommodation or distraction strategies when dealing with problems that are uncontrollable (e.g., medical issues or parental problems) (Compas et al., 1991; Zimmer-Gembeck et al., 2011). This finding suggests that, although the repertoire of coping strategies is still expanding at this age, coping can also become less flexible in its execution during adolescence as youth discern the most effective strategies for dealing with common problems and begin to deploy the strategies more efficiently and with less effort. Throughout adolescence, social partners continue to be crucial to the development of adaptive coping systems. Close and caring parental, family, and peer relationships along with meaningful cultural roles and activities (e.g., after-school programs, sports teams, youth groups) provide the external regulatory structures needed to keep adolescents safe during this risky period while their own coping capacities are still developing. Judicious support from caring adults (parents, teachers, extended family, etc.) is central, in which these grown-ups carefully monitor adolescents' “adventures in coping” (even when adolescents are not keen on having their activities monitored) while remaining sympathetically available to be called on as a backup system for whatever advice and comfort might be requested. Since adolescents' willingness to reveal their problems and failures to adults is a sensitive proposition at this age, the past dyadic history of communication, comfort, and joint interpersonal coping is critical in determining whether, when feeling overwhelmed, adolescents will bring their current coping concerns to an adult. Reflection and Reappraisal Adolescents' growing capacity to reflect on the entire coping system expands their potential to view and utilize coping episodes, whether they are “successes” or “failures,” as increasingly valuable opportunities for learning. Reappraisal processes following mistakes and failures, which were practiced in cooperation with adults during middle childhood, now become highly

inferential and can be reconstructed on the plane of the adolescents' own reflective coping systems, providing internal metaregulatory and metamotivational “commentary” during each phase. Supportive adults, through caring, coaching, and modeling, can help youth cultivate compassion for the fallibility of the self and others in stressful circumstances while also holding the self and others accountable for their actions, thus allowing adolescents to increasingly take ownership for the ways they deal with challenges and setbacks (i.e., for their own coping). These experiences can lead adolescents to supplement their coping repertoires with proactive or antecedent regulation or coping (Aspinwall & Taylor, 1997; Gross, 1998; Gross & Thompson, 2007), in which they make intentional decisions about the states or situations they let themselves get into, to make sure that their regulatory and coping capacities allow them to create the outcomes that they desire.

Conclusion At each level, ways of coping are supplemented with new means of reactivity and regulation. Strategies that were initially neurophysiological and external at birth become interpersonal during infancy and toddlerhood, then individual or intrapersonal during early childhood, eventually to be reconstructed at the reflective level during middle childhood and finally transformed into metacognitive proactive strategies during adolescence and emerging adulthood. At each age, previous means of coping are differentiated and integrated with new modes, supplementing them and providing a wider repertoire that can be deployed more intentionally and flexibly, in concert with changing internal and external conditions and resources. Taken together, these developmental shifts can be seen as creating a system that gets better and better at detecting and responding to threats and problems, eventually realizing its full developmental potential. As we tried to make clear in our earlier work (Skinner & ZimmerGembeck, 2007): These developmental potentials depict a system that can increasingly monitor and appropriately appraise more (current and future) demands using its own and other's “radar”; maintain composure under higher levels of appraised threat with more capacity to withstand multiple demands and better “fallbacks”; respond increasingly in measured socially competent ways that reflect integration of ongoing emotional, attentional, and motivational reactions; more flexibly adjust actions to meet changing environmental demands without losing sight of genuine priorities; recover more quickly from setbacks; and at the same time take more away from stressful encounters, learning how to prevent and deal with future challenges and how to deploy coping in line with future goals. (p. 136)

Normative Development of Coping and Developmental Psychopathology An understanding of the normative reorganizations of the coping system is helpful to researchers interested in developmental psychopathology for multiple interrelated reasons

(Cicchetti & Toth, 2009; Egeland, 2007; Rutter, 2005; Sroufe, 2007). Most important, it allows researchers and interventionists to fully appreciate the complexity of the coping system and the many essential elements orchestrated and consolidated within it. From a developmental perspective, appraisals are not simply fleeting perceptions, and ways of coping are not simply lists of strategies; appraisals reflect “apparent reality” (Fridja, 1987), and ways of coping reflect basic configurations that infants, children, and adolescents have adopted as they organize their actions to deal with demands in the environments in which they are developing (Bowlby, 1969/1982; Garmezy & Rutter, 1983; White, 1974). A normative description provides a basic yardstick for determining whether children and adolescents of all different ages are on track in the development of their coping resources and strategies (i.e., whether they seem to be accomplishing the age-graded developmental tasks that would serve the dual purpose of allowing them to deal effectively with current demands and stressors as well as providing a foundation that prepares them to deal with subsequent challenges successfully). It also allows interventionists to calibrate their prevention and remediation efforts to target the attitudes and skills that are most central to children at their current developmental levels as well as those they will need to be successful at subsequent ages. Explanatory theories and research depict the many factors that allow the healthy development of the coping system to proceed and so allow researchers and interventionists to identify positive pathways through which children and adolescents can obtain robust tools for dealing constructively with stress. They also allow researchers to identify children and adolescents who may benefit most from interventions, because a range of problems in the subsystems that underlie or scaffold coping may put children at risk for the development of psychological and behavioral difficulties. Prevention and remediation efforts could rely on explanatory research to help locate likely intervention levers that would allow the systems that support the healthy development of coping to be repaired or optimized. At the broadest level, developmental conceptualizations have created much overlap between the study of coping and the study of the development of psychopathology and resilience. Approaches that frame coping as part of multilevel integrative systems designed to detect and respond to threats and challenges have identified subsystems that extend from neurophysiology to social context and culture. Such perspectives on the development of coping create broad conceptual overlap with the approaches that have guided the study of developmental psychopathology (Cicchetti & Toth, 2009) and resilience (Compas, 2004; Masten, 2007) over the last five decades (Luthar, 2006). All these traditions consider their phenomena to extend from the biological (genetic, neurophysiological) through the behavioral to the psychological, social, and cultural levels of analysis (Cicchetti & Curtis, 2007). Definitions of coping as “reactivity and regulation under stress” link coping to the burgeoning research on regulation of the many processes (e.g., emotion, attention, behavior, motivation) that are activated during stressful encounters and that have been found to be implicated in the development of psychopathology (Cicchetti & Toth, 2009). To understand coping—and especially its development—it is necessary to understand the

multiple systems that give rise to it and how they work together over time. This is a central tenet of an integrative multilevel approach. The assertion that coping actions are diagnostic of the state of the entire “coping system,” which includes psychological and social stressors, demands, resources, and their underlying neurophysiological and overarching social contexts, suggests that during transactions with stress, children and adolescents can and do move through successive (re)appraisals of challenge, threat, and loss as well as through a repertoire of adaptive and maladaptive ways of coping. Hence, when current developmental researchers refer to the “coping system,” they typically view coping actions as outward manifestations of an orchestra of biopsychosocial forces, very similar to the ways resilience researchers refer to their phenomena, mutatis mutandis. In other words, coping is part of an open and dynamic system and, as such, can be considered part of the same set of (sub)systems that contribute to (and are reciprocally shaped by) developing psychopathology and resilience (Marshall, 2013).

Developmental Systems Perspectives: Coping as Part of Developmental Cascades toward Psychopathology and Resilience A description of normative age-graded reorganizations of the coping system provides an outline of a set of broad and healthy pathways that lead toward constructive and cooperative repertoires for detecting, dealing with, and learning from challenges, threats, and losses. The identification of the many subsystems that work together to give rise to coping actions, as depicted in Figure 10.4, makes it clear that optimal development and the avoidance of psychopathology entail a large number of essential ingredients, which typically serve as reciprocally self-righting, in that they work together to assist the organism toward pathways that channel growth in positive directions. For example, the newborn's intact neurophysiological and communication subsystems, the caregiver's sensitive responsiveness, and their emerging joint secure attachment, co-create a calmer and more stable infant stress reactivity subsystem that is easier to soothe and also provides clearer communication to the caregiver, which in turn informs and sustains sensitive responsive caregiving and lays down implicit infant learning that supports benign appraisals of ambiguous interactions and triggers constructive coping through focused exploration and proximity seeking. At the same time, if any of these essential ingredients is missing or incapacitated, it can pose a risk for children and adolescents, nudging them toward maladaptive coping and psychopathology. In principle, problems can arise from any of the components described in the previous section on normative development, but we focus on three factors that play crucial roles in shaping the architecture of the coping system: (1) differences in the stress and regulatory neurophysiology that underlie coping, especially as captured in work on temperament; (2) differential histories of caregiving relationships involved in scaffolding action tendencies and regulatory processes under stress, especially as examined in work on attachment and parenting; and (3) differential exposure to stressful life events (especially those that originate in the family) that can undermine or overwhelm developing coping systems. To integrate and extend research that has documented links among these factors, maladaptive coping, and psychopathology, we use several key principles from dynamics systems perspectives, which have been clearly articulated in approaches to developmental psychopathology (Cicchetti & Toth, 2009; Masten, 2007; Sroufe, 2009). First, building out from previous research, we assume that the connections between coping and psychopathology are bidirectional and cascade over time. Second, we explore the idea that one way in which risk factors can exert their effects on psychopathology is by creating perturbations in (neurophysiological, interpersonal, individual, reflective, or proactive) coping systems. These perturbations should create lasting effects to the extent that they not only interfere with the concurrent functioning of the coping system but also damage the foundations that are under construction and upon which subsequent developments of the coping system will be built. Third, we think it unlikely that any one factor (e.g., difficult temperament or avoidant attachment) would be sufficient by itself to eventuate in a nonnormative pathway of

maladaptive coping. In cases where a single factor is a little off, protective and compensatory processes can (and typically do) redirect divergent pathways back toward healthy development. Instead, risk accrues when the functioning of one factor is very seriously compromised or is part of a pattern of cumulative impairment or stress, because these conditions may combine to initiate self-amplifying patterns of maladaptive (neurophysiological, interpersonal, and individual) coping. Fourth, there are likely to be multiple pathways from coping to psychopathology, and different families of coping, rather than being considered global indicators of “good news” or “bad news,” might suggest different markers of and routes toward different disorders. Finally, episodes of maladaptive coping may be expected to routinely participate in a wide range of “developmental cascades” (Masten & Cicchetti, 2010), both signaling and igniting problems in multiple domains—problems in individual functioning and in relationships—that cumulatively put children and adolescents at risk for the development of psychopathology. In this section, we select illustrations of how extremes in the functioning of each of these three subsystems (i.e., temperament, attachment and parenting, and family stress) can act as risk factors for both coping and psychopathology. Whenever possible, we focus especially on longitudinal studies that directly explore this possibility (for more detailed analyses, see Skinner & Zimmer-Gembeck, in press). In every case, the research base is inconsistent and cannot confirm all the connections that we posit. In fact, consistent with previous descriptions of the coping literature as dominated by correlational studies of individual differences in ways of coping, much of the research connecting coping to temperament, attachment and parenting, and family stress is cast in this mold, making connecting the developmental dots challenging. At the same time, however, when the findings are considered as a whole, we think they are encouraging and together suggest that it may be useful to further explore the ideas presented in this section. Specifically, we argue that coping is always a part of the developmental cascades that lead to psychopathology: The neurophysiological, psychological, and social factors that predispose children to the eventual development of psychopathology typically result in maladaptive forms of stress reactivity and coping, and these forms of stress responses in turn set successive chains of negative events in motion, including the formation of habits, action tendencies, implicit and explicit self-systems, and (perhaps especially) reactions from social partners that are likely to further exacerbate behavioral problems and cumulatively potentiate the onset or escalation of disorders. These ideas are depicted graphically in Figure 10.5, which shows the cascades that start with temperament and attachment, are embedded in family stress, and, along with parenting, lead to different forms of psychopathology or resilience.

Figure 10.5 Underlying Neurophysiological Factors and Overarching Socialization Factors that Contribute to the Differential Development of Maladaptive Coping and Increase the Risk of Behavior Problems and Psychopathology.

Temperament, Differential Pathways of Maladaptive Coping, and Psychopathology Researchers have suggested multiple ways in which temperament and personality factors contribute both directly and indirectly to the development of problem behaviors and psychopathology (e.g., Calkins, 1994; Carver & Connor-Smith, 2010; Caspi, Roberts, & Shiner, 2003; Connor-Smith & Flachsbart, 2007; Derryberry et al., 2003; Mezulis, Hyde, & Abramson, 2006; Nigg, 2006; Rothbart, 2011, Table 10.1; Rueda & Rothbart, 2009; D. Watson, Kotov, & Gamez, 2006). For example, research indicates that children who have an easy temperament are less prone to develop both internalizing (e.g., depression and anxiety) and externalizing problems (e.g., behavioral problems; Jaffee, Caspi, Moffitt, Polo-Tom, & Taylor, 2007), and children with temperaments characterized as high in negative reactivity or as very inhibited may be at risk for anxiety and other disorders (Belsky & Pluess, 2009; Eisenberg et al., 1997; McClure & Pine, 2006). This body of research is substantial and complex, and has been very useful for identifying individual differences in the kinds of physiologically based temperamental characteristics that may tip a child toward or away from developmental

pathways marked by maladaptation or health and resilience. Yet it may also be useful to supplement this growing body of research by explicitly considering the differential patterns of stress reactivity and coping that stem from such temperamental predispositions in order to understand when, how, and why temperament provokes risk or provides protection from psychopathology. Temperamental Patterns as Differentially Tuned Primitive Coping Systems In principle, the coping systems of all newborns are built on their own individual temperaments, which refer to heritably based neurophysiological processes that predispose them to patterns of responding to environmental challenges and stressors (Rothbart, 2011). Most interesting from a coping perspective are temperamental classifications that focus on reactivity and regulation (Derryberry et al., 2003; Rothbart, 2011; Rueda & Rothbart, 2009; Shannon, Beauchaine, Brenner, Neuhaus, & Gatzke-Kopp, 2007). In this context, “reactivity” is closely connected to stress reactivity, in that it refers to how easily the appetitive/approach and the defensive/ inhibitory systems can be triggered by external and internal stimuli; and “regulation” has close ties to action regulation, in that it refers to how effective the executive attention system is in facilitating volitional control of emotional, motor, and attentional reactivity (Rueda & Rothbart, 2009). As pointed out by Derryberry et al. (2003), “the appetitive and defensive systems can be viewed as relatively primitive ‘coping’ systems. The defensive system is designed to help the person cope with dangerous situations where it is crucial to recognize the threat, inhibit inappropriate responses, and find a source of safety. In contrast, the appetitive system is designed to help the person attain positive outcomes in appetitive contexts, where it is crucial to avoid or overcome obstacles in order to obtain the reward” (p. 1052). Individual differences in newborns' stress responses and regulation (i.e., coping) are shaped by the balance among these three neurophysiological systems: the appetitive, defensive, and executive attention systems. The appetitive and defensive systems are generally antagonistic in their functioning, so that overactivation of one system can be modulated by the activation of the other. Moreover, over- or underactivation of the appetitive or defensive systems can be compensated for by the executive attention system, which, if capable enough, can modulate the effects of the other two systems. Research on Temperament and Coping A growing body of research has focused on direct associations between coping and temperament or personality during childhood and adolescence by examining whether particular coping responses are more or less common in children and adolescents with particular temperamental characteristics or personality traits, such as negative emotionality, neuroticism, effortful control, or introversion (Markovic, Rose-Krasnor, & Coplan, 2013). Several of these studies show, for example, that an “easy” temperament is associated with constructive ways of coping, such as problem solving and support seeking (Zimmer-Gembeck et al., 2011), and with resilience (Luthar, 2006; Werner, 1993). Taken together with research on the links between temperament and psychopathology, such studies provide hints about how temperament might

shape coping under conditions of stress. Most interesting are studies that have begun to more directly investigate these processes by focusing on whether coping serves as a moderator or mediator in the connections between temperament and psychopathology. Moderation has been found in studies testing whether temperamental traits are more relevant to the development of symptoms of depression or anxiety (or other forms of psychopathology) depending on the presence or absence of certain coping responses (Seiffge-Krenke, 2011; Sugimura, Rudolph, & Agoston, 2014). In general, these studies focus almost exclusively on the more risk-producing temperamental traits, such as negative emotionality, and yield patterns of findings suggesting that adaptive forms of coping (such as reliance on primary control strategies when stressors are controllable and on secondary control or accommodative coping when they are not), can reduce the elevated or increasing levels of internalizing and externalizing symptoms that children or adolescents with these temperamental characteristics would otherwise be likely to experience over time. Mediation has been found in studies testing models in which coping acted as a conduit of the impact of particular temperament or personality traits on symptoms or healthy outcomes (Compas, Connor-Smith, & Jaser, 2004; Mezulis, Simonson, McCauley, & Vander Stoep, 2011; K. S. Miller et al., 2009; Van De Ven & Engels, 2011). For example, in one longitudinal study of preadolescents, avoidant coping mediated the impact of temperamental impulsivity on heightened internalizing problems. Thus, preadolescents who were more impulsive used more avoidant coping, which in turn was related to their increasing internalizing symptoms over 1 year (Thompson et al., 2014). Some studies have examined both moderation and mediation, showing that they are not mutually exclusive; they can co-occur even when the same coping constructs, temperamental traits, and adjustment outcomes are examined (K. S. Miller et al., 2009). “Easy” and “Difficult” Temperaments When the research on temperament is filtered through the lens of developmental perspectives on coping, it suggests that temperamental characteristics of newborns can be thought of as neurophysiological set points for the development of their coping systems and so provide early foundations that start to channel infants' coping down differential pathways. The healthy development of coping should be facilitated by an “easy” temperament, which involves moderate responsiveness of the appetitive and defensive systems along with the capacity to flexibly modulate emotion and action using attentional and behavioral processes (Rothbart, 2011). Moderate stress reactivity comprises a threat detection system that is calibrated appropriately to external demands and conditions, in both approach and defense, along with the capacity to flexibly modulate reactions when conditions improve, allowing rapid recovery from stress. From the first days of life, this kind of temperament should make it easier for newborns to participate in effective interpersonal coping: Distress signals are moderate and informative and so more easily converted into directed distress communications, and infants are more easily satisfied and soothed by caregivers' coping efforts on their behalf (Rothbart, 2011). More “difficult” temperaments, which involve high reactivity and poor regulation, should be a

liability in the healthy development of the coping system (see Rothbart, 2011; Rothbart, Posner, & Kieras, 2006). The kinds of problems presented to the developing coping system and their likely effects on subsequent patterns of coping should differ depending on whether overreactivity originates in the defensive/inhibitory or in the appetitive/approach system. In either case, a temperamental profile would be more extreme to the extent that one of the systems is set very high while at the same time the other subsystems that would typically balance it out (namely, the opposing subsystem and the executive attention system) are set very low. Such extreme temperaments create challenges for both infant and caregiver in establishing a calm, well-calibrated and modulated neurophysiological coping system as well as coconstructing a positive interpersonal coping system. Based on their low levels of effortful control, such temperamental patterns may also interfere with the shift toward voluntary selfregulated coping and the subsequent systems that build on this type of coping (i.e., the reflective and proactive coping systems). Inhibited “Fearful” Temperaments, the Differential Development of Maladaptive Coping, and Internalizing Psychopathology Most research attention has been paid to infants with inhibited fearful temperaments that likely reflect an overreactive defensive/inhibitory system (N. A. Fox, Henderson, Marshall, Nichols, & Ghera, 2005; Kagan, 1997; Kemen & Block, 1998). By definition, newborns high in neurophysiological stress reactivity consistently overreact to external demands and internal states. In response to novel stimulation, they typically show high and escalating emotional distress, in which attention is captured by the eliciting event (e.g., a noise or sudden movement). It is as if one of the key underlying systems upon which subsequent coping is going to be built has a hair trigger that keeps tripping the alarm of threat detection constantly and unnecessarily, making such infants' harsh distress signals less discriminating and therefore harder for caregivers to read and interpret. Moreover, weak executive attention (or effortful control) makes it more difficult for infants or caregivers to “turn off” the alarm through soothing, distraction, or other means of interpersonal coping. Attention is not easily freed from the distressing stimulus, and recovery from distress is slow and fragile, with increased risk of triggering additional distressed emotional reactions. High levels of stress reactivity combined with low levels of effortful control may put inhibited/fearful infants and young children at risk for the development of poor coping. An overreactive defensive/inhibitory system provides a more challenging stress neurophysiology, at the same time that it makes sensitive responsive caregiving more difficult, thus interfering with the kind of secure attachment most needed to contribute to the development of a calmer hyporesponsive neurophysiological system (Nachmias et al., 1996). If neither infants nor caregivers are successful in helping recalibrate an overreactive stress neurophysiology, infants and then young children will have repeated experiences of being overwhelmed by internal and external events. Such experiences may in turn lead to overreliance on maladaptive forms of coping, in which children either try to avoid contact with stressors and/or with their overpowering emotional and physiological reactions to them or simply become resigned and submit to overwhelming

stress. Such coping episodes can cumulatively lead young children to construct internal working models that consolidate the “message” of the overreactive defensive system, namely, that the world is a highly dangerous and uncontrollable place with which the self cannot hope to cope effectively. Over time, such a pattern of maladaptive appraisals and ways of coping might mark one step along the pathway to social avoidance, learned helplessness, rumination, and subsequent internalizing problems of depression and anxiety (Derryberry et al., 2003; Keiley, Lofthouse, Bates, Dodge, & Pettit, 2003; Lengua, Sandler, West, Wolchik, & Curran, 1999; Rothbart, 2011). Impulsive “Fearless” Temperaments, the Differential Development of Maladaptive Coping, and Externalizing Psychopathology Although exuberance and “surgency” are typically seen as parts of an “easy” temperament, in some cases, a high appetitive motivational system, unchecked by the defensive system or by effortful control, can still be considered a risk factor for the development of maladaptive coping (Derryberry et al., 2003). An “overactive” appetitive system can lead to forceful and impulsive actions, in which infants and young children fearlessly go after whatever it is they want without attending to dangers or to caregiver attempts to redirect their efforts. Such infants should show a pattern of strong approach tendencies, determined attempts to overcome whatever they perceive as obstacles, persistence in the face of attempts to distract or redirect them, and high levels of frustration and protest if they are thwarted. It is easy for caregivers to see these children, whose every goal becomes a fervent demand, as strong-willed or stubborn, and they can be difficult social partners with whom to construct adaptive interpersonal coping systems because compromise, accommodation, or distraction are not natural action tendencies for them. The lack of discriminating signals about the infants' priorities can make it difficult for caregivers to remain sensitively responsive, and exhausted caregivers can easily decide to just give in and allow infants to have whatever they want, or become frustrated by infants' demands and refuse to cooperate. Both of these reactions, however, create a less than optimal interpersonal coping system. When caregivers simply give in, this reinforces children for their dogged persistence, making them even more tenacious. However, arbitrarily terminating goal pursuit can also trigger the highly reactive appetitive system, which is focused on overcoming obstacles, thereby escalating protest, frustration, and externalizing problems (Bates, Pettit, Dodge, & Ridge, 1998; Shaw et al., 1998). Both of these caregiver reactions may also contribute to experiences that reinforce implicit appraisals of the world as one that requires strong and unyielding opposition on the part of infants if they are to reach their goals. Such generalized expectancies can, in the face of even mild stressors, such as being asked to wait or settle for another goal, cumulatively strengthen prepotent bottom-up action tendencies of explosive reactance or resistance. These reactions may then become increasingly difficult for caregivers to modulate in interpersonal coping episodes and can become very problematic for toddlers and young children to manage with their emerging executive functioning skills, as they encounter the developmental task of selfregulation, a task that is a prerequisite for the construction of the intrapersonal coping system. Without these personal and interpersonal buffers, such impulsive and strong-willed

temperaments may contribute to the development of maladaptive forms of coping, such as opposition or perseveration, that are precursors or markers of emerging externalizing disorders, such as aggression (Eisenberg et al., 1996, 2001; Rothbart et al., 1994). Temperament and Differential Pathways of Coping and Psychopathology Although research on the role of coping in the connection between temperamental characteristics and the development of psychopathology is just beginning, it does seem to suggest at least three important working hypotheses that could be tested in future studies. First, it seems likely that extreme neurophysiological temperamental patterns (which combine high levels of reactivity in the defensive or appetitive subsystem with low levels in both the opposing subsystem and executive attention) should predispose infants and young children to corresponding differential patterns of stress reactivity and coping, and such patterns are likely to be exacerbated by experiences of stress. Second, it seems possible that these strong reactions to stress could be overwhelming to infants and their caregivers and that if caregivers cannot gently curb and compensate for them during infancy, an integrated base of neurophysiological, implicit learning, and relationship problems, as manifest in patterns of maladaptive coping, may emerge as a result. Third, it also seems possible (based on studies with older children) that adaptive ways of coping, first laid down in interpersonal relationships with caregivers, could be one set of strategies that eventually allow young people to adapt and constructively manage the demands created by their own temperaments. By learning to “listen” compassionately to one's own neurophysiological messages about stress reactivity and to respond to them with caring suggestions about how to “keep calm and carry on,” infants and their caregivers may be able to strengthen the (interpersonal and then intrapersonal) coping and regulatory systems that can complement and compensate for the potential neurophysiological liabilities implied by temperamental risk.

Attachment, Differential Pathways of Maladaptive Coping, and Psychopathology Although most attachment theorists agree that patterns of caregiver-infant interactions and attachment relationships formed in the earliest years of life will set in motion differential patterns of coping with stress, empirical research has been slow to directly substantiate these claims. Very few studies have directly examined secure and insecure attachment relationship classifications as correlates of coping responses, and those that do tend to focus on older children. For example, several studies have shown that parental warmth and support, or attachment quality (as measured, e.g., by the Parent and Peer Attachment measure; Armsden & Greenberg, 1987) are associated with more adaptive coping, such as active problem solving or support seeking, in children and adolescents (e.g., Dusek & Danko, 1994; Gaylord-Harden, Taylor, Campbell, Kesselring, & Grant, 2009; Kliewer et al., 1996). Moreover, in one study of coping and emotion regulation in 87 children ages 10 to 12, attachment security was associated with less difficulty recognizing emotions, but it was not associated with more adaptive coping (Brumariu et al., 2012). Instead, it was disorganized attachment, and not anxious or avoidant attachment, that was associated with less active coping and more catastrophizing, suggesting

that children who have experienced some of the most significant caregiver-infant relationship failures will register the greatest impact on their coping (Kobak et al., 2006). Attachment and the Differential Development of Coping When viewed through the lens described in the previous section on the normative development of constructive coping, however, it becomes clear that the healthy growth of the coping system is predicated on sensitive responsive caregiving and a secure caregiver-infant attachment relationship. Initially, sensitive responsiveness provides “external” coping on newborns' behalf, and, cumulatively, it shapes interactions that allow caregivers to co-construct a cooperative interpersonal coping system with infants (Contreras & Kerns, 2000; Raby et al., 2012; Sroufe, Egeland, Carlson, & Collins, 2005). If infants' coping systems have to operate without the scaffold of this sensitivity, and insecure attachment relationships form, infants are required to adapt to caregiving that does not always protect them or help to soothe them effectively and to social interactions that are not tuned to their signals, are not responsive to their expressed needs, and may even add to their distress (Cassidy & Berlin, 1994; Mikulincer & Florian, 2003; Thompson & Meyer, 2007). It is as if, when infants' systems are calibrated to insecure or disorganized attachment relationships, the coping system has received the message that the world is stressful and dangerous, or at least cannot be relied on to provide consistent help or comfort when stress and distress occurs (Kobak, Little, Race, & Acosta, 2001). For infants with highly reactive temperaments, such experiences seem to result in continued stress reactivity, instead of the more normative establishment of equanimity and hyporesponsivity seen in infants who are part of secure attachment relationships (Nachmias et al., 1996). When infants adapt their behavior and coping to these experiences of maladaptive interpersonal coping over the first year, nascent problems can be observed in their physiological and emotional reactivity, communication, and action tendencies, presaging the development of maladaptive patterns of coping with stress (Bosquet & Egeland, 2006; Cicchetti & Rogosch, 2009; L. M. Diamond & Aspinwall, 2003; Mikulincer & Florian, 2003). The particular form that problems with coping will take likely depends on the kinds of responsiveness that is provided, as depicted in different kinds of insecure or disorganized attachments. Anxious-Resistant Attachment Relationships When the early caregiving environment is behaviorally inconsistent and emotionally unreliable —when caregivers are sometimes forthcoming but mostly are unresponsive to expressed needs, and when they can be at times neglectful and at other times intrusive and overstimulating —these experiences tend to yield insecure attachment relationships that are referred to as anxious and/or resistant. When considering the effects of a history of these experiences on coping, it is as if infants must adapt their systems to compensate for the lack of a reliable coping partner by trying to amplify their own part in the exchange, essentially ramping up the volume—leading to greater emotional reactivity, louder and more unrestrained signaling of distress that is harder to turn off, more indiscriminate proximity seeking, and more wary monitoring of the caregivers' whereabouts (Nolte, Guiney, Fonagy, Mayes, & Luyten, 2011).

Moreover, when coping develops within such a shaky interpersonal scaffold, secondary attachment patterns, called hyperactivating regulatory strategies (Mikulincer & Florian, 2003; Wei, Heppner, & Mallinckrodt, 2003), may form and emerge as anxious preoccupation about the availability of support. Thus, the developing organism ends up with little opportunity and energy to devote to learning constructive strategies for exploring the world or dealing with negative emotions or problems. Hence, infants (and later toddlers) are likely to overrely on proximity seeking (which can become dependency), but since the relief that proximity is supposed to provide is available only intermittently, their coping is also laced with irritation, frustration, and opposition, directed toward the caregiver. Because of the lack of contingency between their actions and caregivers' responses, they should also increasingly show coping that is fragile, helpless, and easily derailed. Just as with the overreactive physiologies characteristic of certain temperaments, this history of caregiving would be another pathway to stress reactivity that is so high that it becomes challenging (for anyone) to successfully (co)regulate it. Insecure Avoidant Attachment Relationships A different pattern of coping would be likely to result from avoidant attachment relationships, in which infants must adapt their coping to caregivers who reject their expressions of distress, who show some hostility in response to signals indicating discomfort, and generally appear to resent infants' bids for attention and help. In these relationships, infants basically discover that, if they want to maintain proximity with their caregivers, they must learn to overregulate (Martins, Soares, Martins, Tereno, & Osorio, 2012): Their expressions of distress fall away when it becomes clear that they can elicit negative reactions; they restrain their appeals to caregivers for regulating their emotions or for instrumental help; over time, even when they are in situations that elicit physiological stress reactions, they do not seem to be calmed by the caregivers' presence. Infants who experience a history of rejecting or unavailable caregiving must construct a coping system without the usual reliance on the interpersonal matrix needed to learn adaptive strategies, and so they remain reliant on the primitive actions they can carry out for themselves. The development of such secondary attachment patterns, referred to as deactivating patterns (Mikulincer & Florian, 2003; Wei et al., 2003), results in a coping system that develops within an intrapersonal bubble, unconnected to the wider world of interpersonal resources that could nurture and protect infants locally and that could guide them to more constructive ways of dealing with problems and emotional distress. As a result, toddlers, young children, and adolescents with an avoidant working model of attachment figures may come to rely on social isolation and have difficulty coordinating their coping with other social partners in the future. Such children would be less likely to express (or understand) their distress, less likely to turn to others when they are upset or in trouble, less able to benefit from support offered by others, and less able to cooperate when others try to participate in their coping efforts. Cumulatively, these experiences would be one pathway toward overreliance on maladaptive coping strategies, such as avoidance, escape, and social isolation.

Disorganized Attachment Relationships The third and more recently identified pattern of insecure attachment has been referred to as disorganized; this form of an attachment relationship is more likely to emerge from a history of maltreatment or other multiproblem family circumstances. In the case of child abuse, caregivers not only fail to protect their infants from danger, they actually augment infants' exposure to stress by neglecting to take care of their basic needs or through abusive interactions (Cicchetti & Rogosch, 2009). Hence, the internal coping systems of these infants, which must try to adapt to a dangerous and maladaptive interpersonal coping system, can become confused, exhausted, and chaotic—in other words disorganized—marked by a fundamental dysregulation of emotion combined with chaotic attempts to regulate it (DeOliveira, Bailey, Moran, & Pederson, 2004); to denote the riskiness inherent in these kinds of attachments, they are sometimes referred to as catastrophizing (Brumariu, Kerns, & Seibert, 2012). Insecure Attachments and Stress Reactivity Over time, stress-inducing interactions with caregivers can have a negative effect on the normative development of both neurological and emotional functioning (Cicchetti & Rogosch, 2009; Herbert et al., 2006; McEwen, 1998; 2004; Sapolsky, 1999). Overall, children in dyads with insecure attachment relationships of any kind, but particularly the disorganized form, show greater risk of atypical cortisol responses to threat (Ahnert, Gunnar, Lamb, & Barthel, 2004; Nachmias et al., 1996; Spangler & Schieche, 1998). Although cortisol reactivity to stress is adaptive in the short term, chronic activation of the HPA axis (the systems involved in rapid threat appraisal and response, including parts of the PFC) has been associated with physical and psychological impairment and neuronal death, resulting in a system primed for stress and unable to regulate (recover from) the stress response once it is activated (Gunnar & Cheatham, 2003; Gunnar & Vasquez, 2006). Additionally, cortisol level has been associated with affective responses and regulation. Children have greater cortisol responses when they have less knowledge of emotion control (Gunnar, Marvinney, Isensee, & Fisch, 1989) and when social support resources are not available (Gunnar, Larson, Hertsgaard, Harris, & Brodersen, 1992). These findings suggest that cortisol is higher when individuals perceive they are unable cope or have few response options when confronted with threat or challenge (Dawson, Hessel, & Frey, 1994; Stansbury & Gunnar, 1994), and this reaction is more likely when there is an insecure attachment history. Cumulatively, chronic stress reactivity can result in disruptions to the typical pattern of cortisol over a day; in humans, glucocorticoids demonstrate a circadian rhythm with the highest levels in the morning on awakening and the lowest levels in the evening prior to onset of sleep (Sapolsky, 1992). A departure from this typical pattern has been found among children neglected or otherwise maltreated by their caregivers. Children with such a history tend to show low levels of cortisol in the early morning and blunted adrenocorticotropic hormone and cortisol responses to stressors (referred to as hypocortisolism). Children placed in foster care because of maltreatment also have atypical cortisol diurnal rhythms compared to children

without such histories (Loman & Gunnar, 2010). Secondary Attachment Strategies and Coping Connections between attachment status and ways of coping have been investigated most thoroughly by researchers who have proposed upward theoretical extensions of attachment theory beyond childhood, focusing on attachment as an affect regulation system during adolescence (e.g., Allen & Manning, 2007; Allen & Miga, 2010; Cassidy & Berlin, 1994; Compas, Worsham, & Ey, 1991) or adulthood (Mikulincer & Florian, 2003; Shaver & Mikulincer, 2002). More specifically, current theoretical views highlight the notion of secondary attachment strategies, which depict configurations of emotion regulatory strategies used by older children, adolescents, and adults when dealing with, managing, or confronting stressful life events. According to this perspective, attachment relationships early in life, sometimes included as one aspect of the overall emotional climate of the family (Morris, Silk, Steinberg, Myers, & Robinson., 2007), result in different patterns of emotion recognition and expression and differential utilization of strategies for dealing with stress that emphasize reliance on others versus self-reliance or social isolation. A few researchers have extended the range of secondary attachment strategies under investigation to consider how attachment could be associated with all the core families of coping—from support seeking and active problem solving to emotion expression, avoidance, distraction, and withdrawal during infancy (Roque, Verissimo, Fernandes, & Rebelo, 2013), adolescence (Gaylord-Harden, Taylor, Campbell, Kesselring, & Grant, 2009), and adulthood (Holmberg, Lomore, Takacs, & Price, 2010; Wei et al., 2003). It may be the case that attachment emerges as a stronger correlate of coping after interpersonal and intrapersonal forms of coping have been fully internalized, which we argue takes place by early adolescence. This internalization may help to explain why studies of adolescents and adults, compared to studies of children, typically find more consistent associations between attachment categories and patterns of coping. For example, in one study of adults, the two types of insecure attachments (avoidant and anxious) were differentially associated with theoretically specified ways of coping: Attachment avoidance, but not attachment anxiety, was associated with less use of social support; and attachment anxiety, but not attachment avoidance, was associated with more use of emotion-focused coping (Holmberg et al., 2010). Attachment and Coping with Interpersonal Stressors Patterns of findings linking attachment to coping during adolescence suggest that the effects of attachment history may be more pronounced when individuals are dealing with interpersonal stressors. For example, two studies have shown that secure attachment is related to more active coping with the interpersonal stressors of conflict and relationship dissolution (Creasey & Hesson-McInnis, 2001; Davis, Shaver, & Vernon, 2003). Moreover, one of the few longitudinal studies of these connections from adolescence to early adulthood also found clearer patterns when examining changes in coping with relationship stressors (SeiffgeKrenke, 2006). Researchers followed 64 girls and 48 boys for 7 years through five waves of

data, collected at ages 14, 15, 16, 17, and 21 years. Associations between perceptions of the severity of relationship problems (parent, peer, partners) and three kinds of coping (active coping, which combined problem solving and support seeking, internal coping, and withdrawal) were examined as correlates of attachment classification (secure, dismissing, or preoccupied, as measured via the Adult Attachment Interview at age 21). As expected, the secure group stood out as utilizing increasingly more active coping responses over time when dealing with stressors in both the parent and the peer domains. Secure group members were more active than the dismissing group in responding to parent stress, more active than the preoccupied group when responding to peer stress, and used less withdrawal in response to all stressors when compared to the preoccupied group. Although not directly examined in this study, the pattern of results for the insecure groups suggests that it is not the level of a particular coping response that may signal problems but rather the combination of different responses. For example, the preoccupied group had an ambivalent style of coping that involved high levels of both active coping and withdrawal, and the dismissing group used internal coping at a level that did not differ from the secure group but, compared to the secure group, also used less problem solving and support seeking (i.e., active coping). Such patterning may also be important in detecting differences based on attachment status when coping sequences are examined. Although not yet investigated in children and adolescents, a study of 75 adults revealed that attachment classification was related to the point in the coping sequence at which participants reported using specific ways of coping, such as social support, distancing, and emotion-focused coping (Holmberg et al., 2010). When coping with major events, those classified as avoidant-dismissing (via the Experiences in Close RelationshipRevised; Fraley, Waller, & Brennan, 2000) reported using distancing to cope earlier in the stressful encounter and resorted to seeking support from a partner only later in the coping episode. In contrast, those classified as anxious-preoccupied reported the use of emotionfocused coping earlier in the coping sequence. Coping as a Mediator of the Effects of Attachment on Psychopathology Many researchers have suggested that coping could be a mediator that links attachment to psychopathology, given evidence that both coping and insecure attachment, particularly disorganized attachment, are associated with the development of multiple forms of psychopathology (Kobak et al., 2006). In piecing this literature together, some longitudinal research has shown that the form of insecure attachment classification (or continuous measures of attachment anxiety and avoidance) partially accounts for different forms of emotion dysregulation, consistent with theory, and emotional dysregulation has been found to mediate associations between attachment and psychopathology (e.g., Kullik & Peterman, 2013; Wei et al., 2003). However, little research has examined these pathways directly. In one of the few such studies, children's constructive coping was found to mediate the association between maternal attachment and peer competence (Contreras, Kerns, Weimer, Gentzler, & Tomich, 2000). In a second study, 515 undergraduate students completed measures of attachment anxiety,

attachment avoidance, a range of symptoms of psychopathology, and perceived ineffective coping, modeled as a latent variable marked by three ways of coping, namely (1) low levels of problem solving, (2) low levels of a reflective style involving planning and systematic responses to stress, and (3) high levels of a suppressive style involving denial and avoidance. Using structural equation modeling, researchers discovered that attachment anxiety and avoidance were each uniquely associated with higher levels of perceived ineffective coping and that perceived ineffective coping and attachment avoidance (but not attachment anxiety) both showed unique concurrent associations with heightened psychological distress. Thus, as had been argued, ineffective coping did mediate, at least partially, the associations between insecure attachment and psychopathology symptoms (see Lopez, Mauricio, Gormley, Simko, & Berger, 2001, for similar findings with attachment orientations and distress in college students). Although few studies targeting attachment assess the full range of responses that can be used to cope with stressful events, it is possible to draw on research that assesses a narrower set of emotion regulation strategies in order to discover why or how different forms of attachment, even during infancy and toddlerhood, have implications for the ways individuals react to and cope with stress. For example, in one longitudinal study of very young children (Gilliom, Shaw, Beck, Schonberg, & Lukon, 2002), infant boys classified as securely attachment (at age 1.5 years) were shown to possess more effective anger regulation skills (i.e., emotion-focused coping) at 3 years of age. In a second (cross-sectional) study, all three attachment classifications during infancy (assessed via the Strange Situation) were shown to correspond concurrently to the differential use of theoretically predicted sets of emotion regulation strategies, including positive and negative social engagement, object use, and self-comforting (Crugnola et al., 2011). As would be expected, infants classified as secure, compared to those who were anxious-resistant or avoidant, used more positive social engagement to regulate their distress. Also consistent with theory, anxious-resistant infants displayed more negative social engagement and less object use to comfort themselves compared to the other two attachment groups. Finally, infants classified as avoidant displayed less use of social engagement strategies (combined positive and negative) compared to both other attachment groups and relied more on use of objects to regulate emotion than did infants in the anxious-resistant group (see also Smith et al., 2006, for another study of infant attachment and emotion regulation). Perhaps most important, upward extensions of attachment theory provide a foundation for proposing that individuals with different types of insecure attachments are likely to use different predominant models for regulating emotions (Brenning & Braet, 2013; Ein-Dor, Mikulincer, & Shaver, 2011; Shaver & Mikulincer, 2007). In particular, individuals with an insecure-anxious attachment relationship history have been found to rely predominantly on hyperactivating strategies, which entail heightened negative emotions in times of stress combined with overly dependent or energetic reliance on promixity and support from others. In contrast, individuals with an insecure-avoidant relationship attachment history seem to use as their predominant approach the deactivation or suppression of social needs and emotions and the inhibition of proximity seeking. Similar ideas have been proposed in self-determination

theory, which identifies emotion dysregulation and suppression as two primary contrasting styles of emotion regulation (Ryan, Deci, Grolnick, & La Guardia, 2006). The literature remains quite limited, but emerging (mostly cross-sectional) evidence supports theoretical ideas regarding the differences in emotional regulatory approaches that would be expected between those classified as insecure-anxious compared to those classified as insecure-avoidant (see e.g., Cassidy & Berlin, 1994; Thompson & Meyer, 2007). In a series of two cross-sectional studies of Belgian adolescents, youth with insecure-anxious attachments showed greater dysregulation of emotion, whereas youth with insecure-avoidant attachment showed more suppression of emotion (Brenning, Soenens, Braet, & Bosmans, 2012). In a second series of two studies, the type of emotion mattered (Brenning & Braet, 2013): Youth with anxious attachments showed greater dysregulation of sadness and anger, whereas youth with avoidant attachments showed more suppression of sadness but more dysregulation of anger. The authors argued that the type of emotion should shape how it is regulated by insecure-avoidant youth—because expressions of sadness elicit support from others whereas expressions of anger warn others to back off. Thus, it makes sense that insecure-avoidant adolescents, who wish to be left alone, should deactivate sadness but amplify anger. Attachment and Differential Pathways of Coping and Psychopathology Taken together, this work suggests that early attachment relationships may set a template for current interpersonal and subsequent intrapersonal strategies for coping and emotion regulation. Research supports the notion that individuals' emotional and behavioral expressions in response to stress, which include emotion regulatory as well as coping responses, are grounded in their attachment relationship history—and so should differ markedly between individuals with a history of secure versus insecure attachment relationships as well as between individuals with the two different insecure types of relationships. This research also suggests three hypotheses that warrant further investigation. First, insecure forms of attachment, which originate in infancy and toddlerhood, likely have quite specific relations with emotional reactions and relate to different coping responses and emotion regulatory processes. These may be adaptive for managing stress in the short term but may limit optimal development and put infants at risk for the development of psychopathology in the longer term. Second, coping seems to be one important mediator of the association between attachment relationships and later resilience or symptoms of psychopathology. Third, it is possible that many of the sequelae of early attachment relationships for the whole range of coping strategies do not emerge until adolescence or later, when the capacities for intrapersonal and reflective coping are fully integrated. Before this time, infants, toddlers, and young children have a more limited range of coping responses (which are often assessed and studied as emotion regulatory strategies) that are carefully scaffolded (and so may be proscribed) by concurrent interpersonal relationships (Zimmer-Gembeck & Skinner, 2011). Although research is still in its infancy, findings point to the role that coping should play in future studies focused on attachment relationships and their links to the development of psychopathology and resilience.

Parenting, Differential Pathways of Maladaptive Coping, and Psychopathology Reviewers of research on coping and emotion regulation have identified a wide variety of pathways through which parents can shape how their offspring interpret and deal with stressful events (Bradley, 2007, Table 2; Kliewer, Sandler, & Wolchik, 1994; Power, 2004, Tables 1, 2, and 3; as well as Calkins & Hill, 2007; Grant et al., 2006; Kochanska & Kim, 2013; Morris et al., 2007; Schwarz, Stutz, & Ledermann, 2012; Skinner & Edge, 2002a; Thompson & Meyer, 2007). In fact, when examined through the lens of normative development described in the previous section, it becomes clear that the participation of responsive adults, in successively age-graded roles, is essential to every step in the development of children's own adaptive coping. Hence, one way to consider the role that parenting plays in the differential development of maladaptive pathways of coping is to scrutinize each of these parental practices and to focus on their specific negative contribution, considering a wide spectrum of parenting behavior—which might range from allowing children to be exposed to overwhelming stress, to modeling self-blame, to suggesting coping strategies that are ineffective (see Skinner & Zimmer-Gembeck, in press, for a more detailed analysis). A second strategy, more consistent with developmental systems perspectives and upward extensions of attachment theory, would be to consider the interpersonal support systems provided by parents as a whole and to try to consider, when these systems are not adequate, what happens to children's own coping systems when they must adapt to these nonoptimal contexts. Particularly useful in such an endeavor are theories of parenting that identify a relatively small set of umbrella dimensions that describe the basic functions of parenting in meeting children's psychological needs (Bradley, 2007, Table 1; Connell & Wellborn, 1990; Deci & Ryan, 1985; Grolnick, 2002; Skinner, Johnson, & Snyder, 2005). The dimensions that theories converge on, depicted in Table 10.2, include parental provision of warmth, structure, and autonomy support. By the same token, parental interactions that undermine children's basic needs, also presented in Table 10.2, include rejection, chaos, and coercion. Although they are referred to using a variety of labels, the study of each of these dimensions has a long history in work on parenting (see Skinner et al., 2005, for a review over the last five decades).

Table 10.2 Six Dimensions of Parenting Dimension Description 1 Warmth: Through caring and affectionate interactions, parents communicate their emotional availability, unconditional love, and positive regard for the child. 2 Structure: Through dependable, reliable, and contingent interactions and routines, parents create a sturdy durable context that children can count on as being organized, predictable, responsive, and available to provide instrumental help when it is required or requested. 3 Autonomy support: Through interactions that are attuned to the child's own genuine desires and best interests, parents express respect, encouragement, deference, and trust in the child's authentic self. 4 Rejection: Parents overtly or covertly express their aversion, repugnance, or dislike for the child through interactions that are hostile, dismissive, derisive, sarcastic, callous, uncaring, or cruel. 5 Chaos: Parents create a context that is unstable, disorganized, and tumultuous through interactions that are erratic, unpredictable, inconsistent, and noncontingent. 6 Coercion: Parents behave in ways that are controlling, pressuring, and disrespectful, either through intimidation, force, demands for obedience, and threats of punishment or through guilt-inducing criticism and threats of love withdrawal. From this broad perspective, there are two primary ways in which parents can undermine the development of healthy coping: through errors of omission (i.e., by not creating an adaptive interpersonal coping system within which children's coping can develop) and through errors of commission (i.e., by dealing with their offspring in times of stress in ways that are actively unproductive or harmful). These pathways, as described in more detail in the remainder of this section, can potentially provide a rudimentary map to a line of next studies that would add to the growing literature on the role of parents in children's coping (Bradley, 2007; Power, 2004), perhaps extending its investigation to consider children's maladaptive coping as a mediator between risky parenting and the development of psychopathology and resilience (Barber & Harmon, 2002; Compas et al., 2010; Cummings & Davies, 1999; Kobak et al., 2006; Repetti, Taylor, & Seeman, 2002; Rutter, 2013). These dimensions of parenting behavior can be seen as interpersonal resources and pressures that parents can either add or subtract during stressful interactions (Skinner & Edge, 2002a) and that create interpersonal coping systems to which children's own coping must adapt (Grant et al., 2006; McCarthy, Lambert, & Seraphine, 2004; Seiffge-Krenke, 2011; Skinner & Edge, 2002a; Valiente, Lemery-Chalfant, & Reiser, 2007). Negative dimensions of parenting permeate the development of coping systems because they increase objective demands, intensify the experience of threat, escalate distress, interfere with the detection and

identification of the actual nature and source of negative emotions and action problems, trigger action tendencies that are compelling yet destructive, divert personal regulatory resources, lead to over- or underregulation of emotion and action, undermine constructive action regulation, and interfere with learning. For example, when a child is dealing with a demanding situation, parental rejection, such as negative responses to emotional displays (Eisenberg, Fabes, Carlo, & Karbon, 1992; S. Jones, Eisenberg, Fabes, & Mackinnon, 2002) or harsh discipline (Grant et al., 2006) generally makes things worse: It adds the stress of parental disapproval and criticism; it communicates that the parental relationship may be at stake, increasing distress and the probability of threat appraisals; it can trigger action tendencies that propel the child away from the parent (e.g., social withdrawal or escape), shutting down communication of feelings and concerns and thus making it more difficult for the parent to diagnose the problem and be supportive (further subtracting social resources); because the child must also regulate his or her distress about the parent's reaction, it reduces the regulatory resources available for dealing with the stressful situation, making maladaptive coping strategies more likely; and it focuses the child on the conditionality of parental regard rather than on what can be learned from the stressful episode (Assor, Roth, & Deci, 2004). Over time, the interpersonal coping systems created by the general absence of parental warmth, structure, or autonomy support as well as the presence of rejection, chaos, or coercion can be problematic, even when such bouts of parenting are intermittent but especially when they are more chronic, pervasive, and severe, as in cases of parental child abuse and neglect (Cicchetti & Rogosch, 2009; Maughan & Cichetti, 2002; Shipman et al., 2007). They are likely to exert a downward pressure on children's own functioning, cumulatively leading to increasingly consolidated appraisals of implicit threat, maladaptive action tendencies, and poor regulation (Eisenberg et al., 1998; Valiente et al., 2007). The effects of children's attempts to adapt to the interpersonal coping systems created by negative interactions with parents should be visible in their own developing coping systems, as expressed through high emotional reactivity and distress, powerful but ultimately unconstructive action tendencies of different stripes (e.g., reactions to parental coercion that are submissive or oppositional), nonautonomous systems for action regulation that may be weak or rigidly overcontrolled, and a profile of coping that is low in adaptive and high in maladaptive strategies, including a noxious admixture of coping, such as helplessness, escape, social isolation, delegation, submission, and opposition. Cumulatively, these experiences (along with social partners' widespread negative reactions to such maladaptive coping actions) should contribute to children's construction of self-systems that confirm a negative view of the self (as unlovable, incompetent, or inauthentic) or the world (as untrustworthy, uncontrollable, or coercive). These stressful parent-child interactions are also likely to further undermine the quality of the dyadic relationship and so contribute to the development of both personal and interpersonal vulnerabilities for coping. Such maladaptive patterns of appraising and dealing with stress, along with the reactions they provoke from social partners (not just from parents, but also from other family members, teachers, and peers), should put children at risk for escalating cycles of stress and ineffectual coping, not only in the family but also in school and with their age mates, perhaps cumulatively

contributing to forms of internalizing and externalizing psychopathologies.

Family Stress, Differential Pathways of Maladaptive Coping, and Psychopathology The family represents a primary source of comfort and support and can serve as a basis for protection during times of stress. For this reason, stressors that occur in the family, and which threaten its connectedness, can have detrimental effects on health and reduce the likelihood of resilient outcomes (Lynch & Cicchetti, 1998; Repetti et al., 2002; Schwarz et al., 2012). Researchers have identified a long list of family-level factors that put children at risk for the development of behavior problems; these include marital discord, conflict or violence, divorce, family turmoil, death of a parent or sibling, parental problems (substance abuse, physical or mental illness, incarceration), high rates of mobility, crowding and noise, and generally living in conditions of poverty and oppression, in a dangerous neighborhood, or in communities exposed to environmental pollution, war, or natural disasters. These circumstances can all be considered “developmentally challenging” because of their effects on children: They “pose direct harm, have the potential to seriously undermine emotional security, dislodge productive coping strategies, and impede the use of existing assets or the formation of new ones” (Bradley, 2007, p. 102). At the same time, these situations are challenging to parents: To offset the potential harm posed by these circumstances, children require more from their parents, while, at the same time, because these conditions have an impact on everyone in the family, parents have fewer resources to employ in the service of supportive parenting. Hence, researchers who study the ways that children cope with family stressors point out that events like divorce or death of a parent or sibling are doubly dangerous to children, because they stress children directly (creating neurophysiological, psychological, and social problems) and because they stress parents—and so disrupt the family system just when children need its support the most (Sandler et al., 2000). Although full coverage of this topic is beyond the scope of this review, researchers also point out that these developmentally challenging circumstances are not randomly distributed throughout the population (Luther, 2006). Instead, they represent profiles of cumulative risk (Sameroff, 2010) that are common to particular societal niches, such as those created by poverty and oppression, which disproportionately contain environmental risk and dangers and lack supportive resources (Evans, 2004; Evans & Kim, 2012). Hence, parents with potentially limited caregiving resources, such as adolescent single parents, are often asked to parent in conditions that are high in demands (e.g., multiple young high-needs children) and stressors (e.g., poverty, violent relationships, dangerous neighborhood) and low in supports (e.g., neglectful extended family, limited social services, few high-quality child care facilities). These niches can be considered high in cumulative risk, or allostatic load, which is a term that refers to the total concentration or aggregate of stressors to which a person (or dyad or family unit) is subjected or exposed (McEwen, 2010). Concepts of cumulative risk and allostatic load draw researchers' attention to the larger context of stressors and supports within which

children and their families function and can help to explain the vulnerabilities in family systems created by living in niches that attract a high allostatic load (C. Blair & Raver, 2012; Seeman, Epel, Gruenewald, Karlamangla, & McEwen, 2010; Tolan & Grant, 2009). Perhaps most important, as outlined in models of how contexts shape stress and coping responses (Tolan & Grant, 2009) and in models of how coping and coping resources influence psychopathology and health (Taylor & Stanton, 2007), these kinds of stressful contexts are one source of the kinds of stressors children experience, the resources that are available to cope with stress, the coping patterns that emerge, and the development of psychopathology or resilience in the face of adversity. Higher-order contexts, such as culture, neighborhood, or society, will contribute to parentchild coping episodes that, over time, enhance or reduce both personal and social regulatory resources. The accumulation of these episodes should play a significant role in the development of self-system processes that consolidate negative views of each other and prevent both parties from learning how to deal constructively and cooperatively with problems and obstacles. In that sense, family stress can cumulatively contribute to the development of adaptive or maladaptive coping in both children and their parents, and coping responses then becomes some of the core foundations for the development of resilience and psychopathology (Cicchetti & Rogosch, 2009; Tolan & Grant, 2009).

Future Research and Translation of Research into Action Consistent with many researchers who study risk and resilience, we view the factors discussed in this chapter as part of developmental cascades (Curtis & Cicchetti, 2003; Masten & Cicchetti, 2010), in which early conditions (both neurophysiological and social) contribute to subsequent difficulties. These difficulties then snowball through the accumulation of risk factors, to eventually potentiate behavioral problems and psychological disorders. According to this perspective, each of the factors discussed in this chapter can be seen as a series of steps along pathways that become more difficult to reverse the longer they are followed. For example, when infants have temperaments that are high in stress reactivity, it is more difficult for caregivers to be sensitive and responsive to them; as a result, parents and children are more likely to form insecure or disorganized attachment relationships. The parent-infant interactions characteristic of these kinds of attachment patterns shape the developing neurophysiological processes underlying reactivity, sensitizing infants and young children to the effects of stress, perhaps even at the epigenetic level (C. Blair & Raver, 2012). Moreover, with highly stress reactive children, it can become increasingly difficult for parents to remain warm, involved, structured, and autonomy supportive, and so early patterns contribute to (or are the beginnings of) more general styles of problematic parenting. The discordant parentchild interactions characteristic of these styles of parenting contribute to an atmosphere of tension and turmoil in the larger family system. This higher-order family stress permeates caregiving and has its biggest effects on children who are biologically more susceptible to

environmental effects (Cicchetti & Curtis, 2006; Cicchetti & Rogosch, 2009; Ellis, Boyce, Belsky, Bakermans-Kranenburg, & Van IJzendoorn, 2011).

The Role of Coping in Developmental Cascades toward Psychopathology and Resilience These general dynamics have been noted by researchers from many areas, and in fact, temperament, attachment, parenting, and family stress can be considered the usual suspects in creating developmental cascades that lead to almost any kind of problem behavior or form of psychopathology (e.g., Dodge, Greenberg, & Malone, 2008). What this chapter has to add to such discussions is the idea that, during infancy, childhood, and adolescence, patterns of maladaptive stress reactivity and coping are important parts of these cascades—both as symptoms and as players. For example, temperamental vulnerabilities, almost by definition, are most apparent under conditions of stress and are typically held to be vulnerabilities precisely because they trigger maladaptive reactions to external stressors (such as changes, novelty, and social or attentional demands). One of the primary things that makes children with “difficult temperaments” so difficult is their poor coping—their overreactions and their difficulties in accommodating to demands and recovering from stress. By the same token, poor coping is a key marker of insecure, especially disorganized, caregiver-child attachments—the term coping could be used to describe the outward behaviors of infants and young children under the stress of separation and especially during recovery from that stress. For young children with a history of insecure attachments, these patterns of emotional reactivity and maladaptive coping, as carried forward into preschool, are one of the main reasons they find it more difficult to form optimally safe, supportive, and stable relationships with other adults (e.g., teachers) and peers (e.g., Contreras et al., 2000; Kobak et al., 2006). The same point can be made for problematic parenting and family stress: These conditions are risk factors for developmental outcomes for many reasons (Repetti et al., 2002), and parenting problems and family stress are conditions that can have deleterious effects on children's and adolescents' coping, including their neurophysiological and psychological stress reactivity and the development of their regulatory capacities for dealing constructively with stress (Bradley, 2007). In future studies, researchers who specialize in either coping or psychopathology may wish to examine these processes explicitly, by focusing on the multiple ways that maladaptive coping directly participates in these kinds of developmental cascades. Investigations can explore whether the emotional reactivity and action tendencies underlying maladaptive coping (e.g., opposition, submission, delegation, helplessness, escape, social isolation) can get children and adolescents in trouble in social contexts that expect mature forms of coping, whether those contexts are organized by teachers, peers, coaches, or friends' families. It is possible that children's maladaptive coping can have the direct effect of repelling positive supports, generating stress, or provoking reactions that are intrusive, retaliatory, or rejecting (Conway et al., 2012; Liu, 2013). Difficult temperaments, high stress reactivity, and insecure attachments all shape development—and these effects may accrue partly because the poor coping they

engender itself elicits negative reactions from adult and peer interaction partners. Researchers who study profiles or patterns of coping may also wish to consider whether maladaptive coping participates in developmental cascades through the downward pressure it exerts on adaptive functioning. Chronically activated threat appraisals and poor coping may interfere with children's access to their own better natures and competencies, such as their compassion and regulatory resources, making it more difficult for them to enact adaptive strategies and potentially leading them to feel powerless in the face of their own urges. Moreover, the emotional and action impulses characteristic of maladaptive coping may be so compelling that they lead children and adolescents to feel that their negative behaviors are fully justified to defend against imminent threat (Lansford, Malone, Dodge, Pettit, & Bates, 2010). As a whole, the recursive components of maladaptive coping systems, with their high stress reactivity, threat appraisals colored by fear or resentment, experiences of ineffectiveness in dealing with problems and emotions, and the negative social reactions they elicit, together may create an integrated biopsychosocial “apparent reality” that should intensify actual and subjective stress and cumulatively channel development toward heightened reactivity, regulatory vulnerabilities, behavior problems, and psychopathology. A comprehensive picture of the developmental cascades that lead to maladaptive coping, and through maladaptive coping to psychopathology, will clearly include more contexts and social partners than we have been able to describe here. Like all other developmental cascades depicted thus far, they will undoubtedly travel through the worlds of school and peers, where children's and adolescents' maladaptive coping will likely put them at risk for academic disaffection and underperformance, peer rejection and membership in deviant peer groups, evasion of supervision by competent adults, and eventual participation in a host of risky activities during adolescence, such as truancy, delinquency, substance abuse, and unprotected sexual activity, that further constrict life paths during emerging adulthood. The researchers who study these phenomena do not label them all as coping or consider the problems they depict as partly the result of maladaptive coping (cf. Spencer, 2006). Nevertheless, coping researchers can profit from what has been learned about the functioning of temperament, attachment, academic progress, peer relationships, teaching, and parenting under stress. And work on coping may also be useful to researchers from these largely disparate traditions by providing some common ground where they can meet to figure out how all these components work together during stressful encounters to shape children's short-term coping and their long-term development, including pathways toward competence and disorder.

Translation of Basic Research on Coping into Action If indeed, as we have argued, coping is both a key marker and a central player in the development of psychopathology and resilience, it follows that work on coping and its development has the potential to provide a platform for building out preventive interventions designed to avert the onset of mental health (or other) problems as well as for creating targeted interventions designed to promote the construction of coping resources, strategies, and efficacy once problems have been identified. Evidence for the utility of coping as an intervention lever

can be found in the substantial portion of the larger stress and coping research agenda that has been dedicated to translating such research into direct action—by developing and evaluating prevention programs implemented in locations with almost universal access to families and/or children (e.g., the schools). Such interventions often rely on psychoeducation to teach coping strategies with the aim of preventing the development of a range of internalizing and externalizing symptoms or of reducing the likelihood of symptom escalation (e.g., P. M. Barrett, Lock, & Farrell, 2005; Eassau, Conradt, Sasgawa, & Ollendick, 2012; J. K. Fox et al., 2012; Frydenberg & Lewis, 2000; Ginsburg, 2009). Other applied stress/coping research has tested programs designed to improve coping resources and reduce distress and symptoms of mental health disorders among select groups who are at risk due to major life stressors, such as parental divorce or serious illness (e.g., Compas et al., 2010; Conrod, Castellanos-Ryan, & Strang, 2010; Sansom-Daly, Peate, Wakefield, Bryant, & Cohn, 2012; Soper, Wolchik, Tein, & Sandler, 2010; Vélez, Wolchik, Tein, & Sandler, 2011). As a whole, these studies have demonstrated the feasibility and efficacy of providing children and adolescents with guided practice in using coping resources or emotion regulation strategies that can be effective in dealing with many stressors or helpful in alleviating distress (Compas et al., 2009, 2010, 2014; Fresco, Mennin, Heimberg, & Ritter, 2013; Kovacs et al., 2006; Suveg, Sood, Comer, & Kendall, 2009; Tein, Sandler, Ayers, & Wolchik, 2006; Weisz, Hawley, & Jensen Doss, 200; Weisz, Thurber, Sweeney, Proffitt, & LeGagnoux, 1997). Other interventions have shown success in increasing coping efficacy (Gonzales et al., 2012; Wolchik et al., 2000). Some of the most interesting studies have examined coping as a mediator between treatment and outcome, documenting that improvement in coping capacity as a result of interventions is an important factor that accounts for whether children will show reduced symptoms of mental health disorders (or other adjustment problems) following treatment (e.g., Compas et al., 2010; Essau, Conradt, Sasgawa, & Ollendick, 2012; Tein, Sandler, MacKinnon, & Wolchik, 2004). Yet despite the growing body of research that examines efforts to directly intervene to improve coping or coping resources or that examines coping as an outcome or as a part of the recovery process, it is surprising to discover just how few studies of this kind have been conducted and that most of them have been completed quite recently. What is equally noteworthy in this literature, however, is the number of studies that have been carried out to examine the efficacy or effectiveness of implementing enhancement programs for youth that include the aim of improving coping but never actually assess it. Nevertheless, these commonly used clinical treatments spotlight the possibility that minimizing the use of maladaptive coping strategies and identifying new adaptive strategies to use instead, through therapeutic conversation, psychoeducation, role play, direct practice, or other techniques, may be effective components of such programs. Programs implementing these methods have been shown to reduce symptoms of mental health disorders, whether the programs are directed toward youth who are experiencing family problems (Garber et al., 2009; Silverman, Kurtines, Jaccard, & Pina, 2009), chronic physical or health conditions (Sansom-Daly et al., 2011), or general and other forms of anxiety and fear (P. M. Barrett et al., 2005; P. M. Barrett, Farrell, Ollendick, & Dadds, 2006; J. K. Fox et al., 2012).

When these efforts are examined in light of research and theory on the development of coping, however, three primary ways are revealed in which the current research agenda on child and adolescent stress and coping could be broadened in order to better integrate basic research with application to clinical and educational practice. First, as can be imagined from our overall emphasis on the development of coping, we would argue that the entire agenda could benefit from more careful and systematic attention to child age as a marker of developmental level. Second, a greater consideration of the roles of child temperament and family relationship history in contributing to coping strengths and vulnerabilities might allow researchers to design more successful intervention strategies by tailoring approaches to match children's specific underlying issues. Third, there could be better integration of measures of coping in translational research, both to identify children and youth who could benefit from interventions and to assess the mechanisms of effects in interventions—via typical questionnaires and via in-the-moment techniques. Child Age and Developmental Level As described in this chapter, the coping system and the multiple individual and interpersonal subsystems that support and influence it show developmental patterns that are linked with age and experience. This fact is recognized implicitly or explicitly in much of the basic research on stress and coping in children and adolescents, as can be seen in age-graded choices about the coping strategies selected for study or, more rarely, in decisions to examine age directly (Zimmer-Gembeck & Skinner, 2011). Such developmental changes can be viewed as obstacles and challenges to treatment research as well, and most interventions have not yet been tested to determine whether they are effective across the age ranges of children who receive the services (Eyberg, Schumann, & Rey, 1998). At the same time, we now have an increasingly rich body of developmental ideas about coping, which have yet to be fully incorporated into the design of prevention and intervention programs to enhance children's coping resources, capacity, and efficacy, to reduce the fallout from stressful experiences, or to otherwise optimize children's functioning in the face of adversity. A lack of integration of research on coping into program development is partly because reviews of basic research (and basic research itself) have not often directly addressed the development of coping with an eye toward guiding intervention. We also suspect that it is partly because few studies that focus on enhancing coping explicitly attend to age either by examining whether age matters to program effectiveness or by using theory and research to intentionally adapt the program to better serve different age groups (and reporting these modifications) (see P. M. Barrett et al., 2005, for an exception; Eyberg et al., 1998). In one recent study (Farrell, Waters, & Zimmer-Gembeck, 2012), when researchers did examine age differences in the cognitive beliefs theoretically linked with anxiety symptoms, such beliefs were found only among adolescents (ages 12–17) and not among children (ages 7–11), consistent with developmental research that reveals age-graded changes in the ability to consider inner thoughts and beliefs during early adolescence and the greater use of cognitive coping responses in adolescents compared to children (see also Essau et al., 2012). Moreover, basic researchers can help interventionists more thoroughly integrate developmental

dimensions in their programs if they themselves attend more closely to developmental theory when designing studies of coping. They should more closely align their selection of age groups with the stressors and coping responses that should be most relevant. For example, enhancing active coping (problem solving or positive cognitive reappraisal) may not be appropriate for all ages or situations. The kind of accommodative coping that is most effective when dealing with uncontrollable stressors likely differs with age—only during later childhood and early adolescence do cognitive reappraisal strategies become accessible, and even then they may require practice if they are to become durable enough to utilize under stress; during middle childhood, distraction may be more effective, and behavioral distraction may be the only option for younger children, who may also need to rely on support and active distraction by others. Similarly, the kinds of coping that is effective with controllable stressors may also require different intervention responses at different ages. Research scrutinizing the kinds of interpersonal relationships that provide developmentally graded support as children begin to enact more self-reliant forms of coping with stress would be helpful for planning interventions directed to children between ages 5 and 7. Research examining how children use different forms of social support, how they balance this with individual efforts, and how this balancing act changes from early to middle childhood would be particularly useful for identifying entry points for ameliorating deficits and bolstering strengths at these ages. And studies focusing on how the range and flexibility of adolescents' strategy use changes as they get older would also be useful for guiding practical work to help young people more intentionally match their coping efforts to situational demands during these age periods. Finally, research using cluster analysis or person-centered approaches (Boxer et al., 2012; Tolan et al., 2002) also suggests that patterns of coping or patterns of other regulatory responses and coping need to be considered as targets for interventions. We know about the “complex ways that coping strategies are interrelated” (Gaylord-Harden et al., 2010, p. 852), and we are beginning to find out that these interrelations may change with age (ZimmerGembeck & Skinner, 2011). As research progresses toward identifying optimal profiles of ways of coping for different developmental periods and the type of stress under consideration (Skinner & Wellborn, 1997), this information would be invaluable for informing practice. If research is carefully designed with clinical implications in mind, future developmental research findings should be much more easily translated into the creation of practical supports, and treatment or prevention programs for young people of all ages. Perhaps coping researchers will be more motivated to explore important developmental issues, such as what kinds of coping are and are not adaptive in particular situations and at particular ages, if they see how important such findings could be to the creation of more effective prevention and intervention practices (see also Gaylord-Harden et al., 2010; Tolan et al., 2002, for further discussions of these issues). Coping, Temperament, and Family Relationship History Universal intervention to assist children when dealing with stress can be beneficial by teaching a range of cognitive or behavioral strategies to put in place and practice before stress occurs

(P. M. Barrett et al., 2006). Nevertheless, selected approaches have the advantage of tailoring intervention techniques—with the goals of compensating for deficits and building on strengths among individual or groups of children identified for inclusion in programs based on history (e.g., parental divorce or maternal depression; Compas et al., 2009, 2010; Soper et al., 2010; Vélez et al., 2011) or personal symptoms, such as anxiety disorders (Kovacs et al., 2006; Waters, Donaldson, & Zimmer-Gembeck, 2008). Such focused tailoring could be guided even more explicitly by studies documenting the different patterns of maladaptive appraisals, stress reactions, and coping strategies children are likely to exhibit, given developmental differences in temperamental status, attachment history, and parenting experiences. This knowledge might also be useful in assisting children to cope more effectively and compassionately with their own temperamental tendencies and to modify their own behaviors with the goal of reducing “dependent” stress events, which are stressful experiences that are partly self-generated (Conway et al., 2012; Liu, 2013). The study of children's differential histories could also be useful in order to identify potential appraisal biases that will prompt inappropriate coping responses (Kochenderfer-Ladd & Skinner, 2002; Zimmer-Gembeck & Nesdale, 2013; Zimmer-Gembeck et al., 2013), including reliance on coping strategies that are not well matched to the situation. Not only could such studies guide interventions as they directly tackle biased appraisals and maladaptive coping, but they could also identify the kinds of promotive opportunities children are likely to have missed, given their temperamental and attachment histories; some children may never have participated in constructive interpersonal coping or experienced practical success in coping, suggesting that interventions could be enriched by providing new opportunities for such children to acquire these skills as well as to build feelings of trust, cooperation, efficacy, and agency (Larson, 2011), which might all feed back into increasingly more beneficial coping patterns over time. Integrating Coping Measures More Fully into Prevention and Intervention Research Better translation of basic coping research into practice also depends on expanding present intervention research to more frequently consider ways of coping, as well as closely related regulatory processes, as “active ingredients in the prevention [or escalation] of mental health problems in children and adolescents” (Compas et al., 2010, p. 623). Coping measures can serve as screening devices to identify children and youth who could benefit from services. Coping also represents a valid and important measure of the goals, aims, and outcomes of treatment (or other types of prevention/intervention programs) for young people and should be assessed more routinely. Given the challenges of measuring coping, doing so will require careful thinking on interventionists' parts. Current multidimensional measures of coping (e.g., Ayers et al., 1996; Connor-Smith et al., 2000), as reported by children, youth, and their adults, provide a valuable menu from which researchers can select the range of ways of coping that are most appropriate to the developmental level of their target population (Zimmer-Gembeck & Skinner, 2011) and best suited to the situation and treatment approach (Eyberg et al., 1998). Measurement development work is needed in order to assess coping in a way that is sensitive to

developmental age, situation, and prior history but is also sensitive enough to detect changes when comparing assessments at pre- and posttreatment or later. Perhaps daily diary assessments of distress and coping (e.g., Walker et al., 2007) will turn out to be especially useful for these purposes. As assessment becomes more sophisticated, such translational research incorporating coping measures will help to locate the particular components of interventions that yield positive and lasting treatment outcomes (Kazdin, 2008).

Summary and Conclusion We hope that some of the ideas outlined in this chapter (such as those summarized in Table 10.1) may provide grist for thought about how future research can productively combine research on the development of coping with work on the development of psychopathology. These are two rich empirical traditions that attempt to understand how adversity and stressful events in the lives of children and adolescents can shape their developmental pathways, for better and for worse. We have suggested that coping—with its strategies, resources, and efficacy—should always be considered a part of these developmental cascades and also a part of intervention and prevention efforts to optimize them. Both developmental psychopathology and the development of coping have much to contribute to our understanding of how children and adolescents, when faced with stressful life events, whether they are woven into lives of privilege or derived from chronic adverse circumstances, can (with or without our help) nevertheless create pathways that lead to the development of enduring competence and resilience.

References Ahnert, L., Gunnar, M. R., Lamb, M. E., & Barthel, M. (2004). Transition to child care: Associations with infant–mother attachment, infant negative emotion, and cortisol elevations. Child Development, 75, 639–650. doi: 10.1111/j.1467–8624.2004.00698.x Ainsworth, M. D. S. (1979). Infant-mother attachment. American Psychologist, 34, 932–937. Aldao, A., & Nolen-Hoeksema, S. (2010). Specificity of cognitive emotion regulation strategies: A transdiagnostic examination. Behaviour Research and Therapy, 48, 974–983. doi: 10.1016/j.brat.2010.06.002. Aldao, A., & Nolen-Hoeksema, S. (2012). When are adaptive strategies most predictive of psychopathology? Journal of Abnormal Psychology, 121, 276–281. doi: 10.1037/a0023598 Aldao, A., Nolen-Hoeksema, S., & Schweizer, S. (2010). Emotion-regulation strategies across psychopathology: A meta-analytic review. Clinical Psychology Review, 30, 217–237. Aldridge, A. A., & Roesch, S. C. (2008). Developing coping typologies of minority adolescents: A latent profile analysis. Journal of Adolescence, 31, 499–517. doi: 10.1016/j.adolescence.2007.08.005

Aldwin, C. M. (2007). Stress, coping and development: An integrative perspective. New York, NY: Guilford Press. Aldwin, C. M., & Revenson, T. A. (1987). Does coping help? A reexamination of the relation between coping and mental health. Journal of Personality and Social Psychology, 53, 337– 348. doi: 10.1037/0022–3514.53.2.337 Aldwin, C. M., Skinner, E. A., Zimmer-Gembeck, M. J., & Taylor, R. (2011). Coping and selfregulation across the lifespan. In K. Fingerman, C. Berg, T. Antonucci, J. Smith, & T. Antonucci (Eds.), Handbook of lifespan development (pp. 563–590). New York, NY: Springer. Allen, J. P., & Manning, N. (2007). From safety to affect regulation: Attachment from the vantage point of adolescence. In M. Scharf & O. Mayseless (Eds.), Attachment in adolescence: Reflections and new angles (pp. 23–40). San Francisco, CA: Jossey-Bass. Allen, J. P., & Miga, E. M. (2010). Attachment in adolescence: A move to the level of emotion regulation. Journal of Social and Personal Relationships, 27, 181–190. Appleton, A. A., Buka, S. L., Loucks, E. B., Gilman, S. E., & Kubzansky, L. D. (2013). Divergent associations of adaptive and maladaptive emotion regulation strategies with inflammation. Health Psychology, 32, 748–756. doi: 10.1037/a0030068 Armsden, G. C., & Greenberg, M. T. (1987). The Inventory of Parent and Peer Attachment: Relationships to well-being in adolescence. Journal of Youth and Adolescence, 16, 427–454. Aspinwall, L. G., & Taylor, S. E. (1997). A stitch in time: Self-regulation and proactive coping. Psychological Bulletin, 121, 417–436. doi: 10.1037/0033–2909.121.3.417 Assor, A., Roth, G., & Deci, E. L. (2004). The emotional costs of parents' conditional regard: A Self-Determination Theory analysis. Journal of Personality, 72, 47–88. Austenfeld, J. L., & Stanton, A. L. (2004). Coping through emotional approach: A new look at emotion, coping, and health-related outcomes. Journal of Personality, 72, 1335–1364. doi: 10.1111/j.1467–6494.2004.00299.x Ayers, T. S., Sandler, I. N., West, S. G., & Roosa, M. W. (1996). A dispositional and situational assessment of children's coping: Testing alternative models of coping. Journal of Personality, 64, 923–958. doi: 10.1111/j.1467–6494.1996.tb00949. Babb, K. A., Levine, L. J., & Arseneault, J. M. (2010). Shifting gears: Coping flexibility in children with and without ADHD. International Journal of Behavioral Development, 34, 10– 23. doi: 10.1177/0165025409345070 Baddeley, A. (1998). Recent developments in working memory. Current Opinion in Neurobiology, 8, 234–238. doi: 10.1016/S0959–4388(98)80145–1. Bagwell, C. L., Schmidt, M. E., Newcomb, A. F., & Bukowski, W. M. (2001). Friendship and

peer rejection as predictors of adult adjustment. New Directions for Child and Adolescent Development, 91, 25–50. doi: 10.1002/cd.4 Band, E. B., & Weisz, J. R. (1990). Developmental differences in primary and secondary control coping and adjustment to juvenile diabetes. Journal of Clinical Child Psychology, 19, 150–158. Barber, B. K., & Harmon, E. L. (2002). Violating the self: Parental psychological control of children and adolescents. In B. K. Barber (Ed.), Intrusive parenting: How psychological control affects children and adolescents (pp. 15–52). Washington, DC: American Psychological Association. Barker, D. B. (2007). Antecedents of stressful experiences: Depressive symptoms, self-esteem, gender, and coping. International Journal of Stress Management, 14, 333–349. doi: 10.1037/1072–5245.14.4.333 Barrett, K. C., & Campos, J. J. (1991). A diacritical function approach to emotions and coping. In E. M. Cummings, A. L. Greene, & K. H. Karraker (Eds.), Life-span developmental psychology: Perspectives on stress and coping (pp. 21–41). Hillsdale, NJ: Erlbaum. Barrett, P. M., Lock, S., & Farrell, L. J. (2005). Developmental differences in universal preventive intervention for child anxiety. Clinical Child Psychology and Psychiatry, 10, 539– 555. Barrett, P. M., Farrell, L. J., Ollendick, T. H., & Dadds, M. (2006). Long-term outcomes of an Australian prevention trial of anxiety and depression symptoms in children and youth: An evaluation of the Friends Program. Journal of Clinical Child and Adolescent Psychology, 35, 403–411. doi: 10.1207/s15374424jccp3503_5 Bates, J. E., Pettit, G. S., Dodge, K. A., & Ridge, B. (1998). Interaction of temperamental resistance to control and restrictive parenting in the development of externalizing behavior. Developmental Psychology, 34, 982–995 Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135, 885–908. Berg, C. A., & Strough, J. (2010). Problem solving across the life span. Handbook of life span psychology (pp. 239–265). Berkel, C., Knight, G. P., Zeiders, K. H., Tein, J.-Y., Roosa, M. W., Gonzales, N. A., & Saenz, D. (2010). Discrimination and adjustment for Mexican American adolescents: A prospective examination of the benefits of culturally related values. Journal of Research on Adolescence, 20(4), 893–915. doi: 10.1111/j.1532–7795.2010.00668.x Best, J. R., & Miller, P. H. (2010). A developmental perspective on executive function. Child Development, 81, 1641–1660.

Blair, C., & Raver, C. C. (2012). Child development in the context of adversity. American Psychologist, 67, 309–318. Blair, K. A., Denham, S. A., Kochanoff, A., & Whipple, B. (2004). Playing it cool: Temperament, emotion regulation, and social behavior in preschoolers. Journal of School Psychology, 42(6), 419–443. doi: 10.1016/j.jsp.2004.10.002 Block, J. H., & Block, J. (1980). The role of ego-control and ego-resiliency in the organization of behavior. In W. A. Collins (Ed.), Development of cognition, affect, and social relations: The Minnesota symposia on child psychology (Vol. 13, pp. 109–102). Hillsdale, NJ: Erlbaum. Bosquet, M. & Egeland, B. (2006). The development and maintenance of anxiety symptoms from infancy through adolescence in a longitudinal sample. Development and Psychopathology, 18, 517–550. Bowlby, J. (1969/1982). Attachment and loss. Vol. 1. New York, NY: Basic Books. Boxer, P., Sloan-Power, E., Mercado, I., & Schappell, A. (2012). Coping with stress, coping with violence: Links to mental health outcomes among at-risk youth. Journal of Psychopathology and Behavioral Assessment, 34(3), 405–414. doi: 10.1007/s10862–012– 9285–6 Bradley, R. H. (2007). Parenting in the breach: How parents help children cope with developmentally challenging circumstances. Parenting: Science and Practice, 7, 99–148. Brandtstädter, J., & Rothermund, K. (2002). The life-source dynamics of goal pursuit and goal adjustment: A two-process framework. Developmental Review, 22, 117–150. Braungart-Rieker, J. M., & Stifter, C. A. (1996). Infants' responses to frustrating situations: Continuity and change in reactivity and regulation. Child Development, 67, 1767–1779. doi: 10.1111/j.1467–8624.1996.tb01826.x Brenning, K. M., & Braet, C. (2013). The emotion regulation model of attachment: An emotion-specific approach. Personal Relationships, 20, 107–123. Brenning, K. M., Soenens, B., Braet, C., & Bosmans, G. (2012). Attachment and depressive symptoms in middle childhood and early adolescence: Testing the validity of the emotion regulation model of attachment. Personal Relationships, 19, 445–464. Bridges L. J. (2003). Coping as an element of developmental well-being. In M. Bornstein, L. Davidson, C. L. Keyes, & K. A. Moore (Eds.), Well-being: Positive development across the life course (pp. 155–166). Mahwah, NJ: Erlbaum. Bridges, L. J., & Grolnick, W. S. (1995). The development of emotional self-regulation in infancy and early childhood. In N. Eisenberg (Ed.), Social development: Vol. 15. Review of personality and social psychology (pp. 185–211). Thousand Oaks, CA: Sage.

Brittian, A. S., Umaña-Taylor, A. J., Lee, R. M., Zamboanga, B. L., Kim, S. Y., Weisskirch, R. S.,… Caraway, S. J. (2013). The moderating role of centrality on associations between ethnic identity affirmation and ethnic minority college students' mental health. Journal of American College Health, 61(3), 133–140. doi: 10.1080/07448481.2013.773904 Bronfenbrenner, U., & Morris, P. A. (2006). The bioecological model of human development. In R. M. Lerner, & W. Damon, (Eds.), Handbook of child psychology (6th ed.): Vol 1, Theoretical models of human development (pp. 793–828). Hoboken, NJ: Wiley. doi: 10.1002/9780470147658.chpsy0114 Brumariu, L. E., Kerns, K. A., & Seibert, A. (2012). Mother-child attachment, emotion regulation, and anxiety symptoms in middle childhood. Personal Relationships, 19, 569–585. Calkins, S. D. (1994). Being alone, playing alone and acting alone: Distinguishing among reticence, and passive- and active-solitude in young children. Child Development, 65, 129– 137. Calkins, S. D., & Hill, A. (2007). Caregiver influences on emerging emotion regulation. In J. J. Gross (Ed.), Handbook of emotion regulation (pp. 229–248). New York, NY: Guilford Press. Carlson, E., & Sroufe, L. A. (1995). The contribution of attachment theory to developmental psychopathology. In D. Cicchetti & D. Cohen (Eds.), Developmental processes and psychopathology: Vol. 1. Theoretical perspectives and methodological approaches (pp. 581– 617). New York, NY: Cambridge University Press. Carver, C. S. (1998). Resilience and thriving: Issues models, and linkages. Journal of Social Issues, 54, 245–266. Carver, C. S., & Connor-Smith, J. (2010). Personality and coping. Annual Review of Psychology, 61, 679–704. Carver, C. S. & Scheier, M. F. (1998). On the self-regulation of behavior. New York, NY: Cambridge University Press. Casey, B. J., Jones, R. M., & Somerville, L. H. (2011). Braking and accelerating of the adolescent brain. Journal of Research on Adolescence, 21, 21–33. Casey, B. J., Tottenham, N., Liston, C., & Durston, S. (2005). Imaging the developing brain: What have we learned about cognitive development? Trends in Cognitive Sciences, 9, 104– 110. doi: 10.1016/j.tics.2005.01.011 Caspi, A., Roberts, B. W., & Shiner, R. L. (2003). Personality development: Stability and change. Annual Review of Psychology, 56, 453–484. Cassidy, J., & Berlin, L. J. (1994). The insecure/ambivalent pattern of attachment: Theory and research. Child Development, 65, 971–991. Causey, D. L., & Dubow, E. F. (1992). Development of a self-report coping measure for

elementary school children. Journal of Clinical Child and Adolescent Psychology, 21, 47–59. doi: 10.1207/s15374424jccp2101_8 Chorpita, B. F., & Barlow, D. H. (1998). The development of anxiety: The role of control in the early environment. Psychological Bulletin, 124, 3–21. doi: 10.1037/0033–2909.124.1.3 Cicchetti, D., & Curtis, W. J. (2006). The developing brain and neural plasticity: Implications for normality, psychopathology, and resilience. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Developmental neuroscience (Vol. 2, 2nd ed., pp. 1–64). Hoboken, NJ: Wiley. Cicchetti, D., & Curtis, W. J. (2007). A multilevel approach to resilience. Development and Psychopathology, 19, 627–955. Cicchetti, D., & Rogosch, F. A. (2009). Adaptive coping under conditions of extreme stress: Multilevel influences on the determinants of resilience in maltreated children. New Directions in Child and Adolescent Development, 2009, 47–59. doi: 10.1002/cd.242 Cicchetti, D., & Toth, S. L. (2009). The past achievements and future promises of developmental psychopathology: The coming of age of a discipline. Journal of Child Psychology and Psychiatry, 50, 16–25. Clarke, A. T. (2006). Coping with interpersonal stress and psychosocial health among children and adolescents: A meta-analysis. Journal of Youth and Adolescence, 35, 10–23. doi: 10.1007/s10964–005–9001-x Compas, B. E. (1987a). Coping with stress during childhood and adolescence. Psychological Bulletin, 101, 393–403. doi: 10.1037/0033–2909.101.3.393 Compas, B. E. (1987b). Stress and life events during childhood and adolescence. Clinical Psychology Review, 7, 275–302. doi: 10.1016/0272–7358(87)90037–7 Compas, B. E. (2004). Processes of risk and resilience during adolescence: Linking contexts and individuals. In R. M. Lerner & L. Steinberg (Eds.), Handbook of adolescent psychology (2nd ed., pp. 263–296. Hoboken, NJ: Wiley. Compas, B. E., Banez, G. A., Malcarne, V., & Worsham, N. (1991). Perceived control and coping with stress: A developmental perspective. Journal of Social Issues, 47, 23–34. doi: 10.1111/j.1540–4560.1991.tb01832.x Compas, B. E., Champion, J. E., Forehand, R., Cole, D. A., Reeslund, K. L., Fear, J.,… Roberts, L. (2010). Coping and parenting: Mediators of 12-month outcomes of a family group cognitive–behavioral preventive intervention with families of depressed parents. Journal of Consulting and Clinical Psychology, 78, 623. Compas, B. E., Connor, J., Osowiecki, D., & Welch, A. (1997). Effortful and involuntary responses to stress: Implications for coping with chronic stress. In B. J. Gottlieb (Ed.), Coping

with chronic stress (pp. 105–130). New York, NY: Springer. Compas, B. E., Connor-Smith, J. K., & Jaser, S. S. (2004). Temperament, stress reactivity, and coping: Implications for depression in childhood and adolescence. Journal of Clinical Child & Adolescent Psychology, 33, 21–31. Compas, B. E., Connor-Smith, J. K., Saltzman, H., Thomsen, A. H., & Wadsworth, M. E. (2001). Coping with stress during childhood and adolescence: Problems, progress, and potential in theory and research. Psychological Bulletin, 127, 87–127. doi: 10.1037/0033– 2909.127.1.87 Compas, B. E., Forehand, R., Keller, G., Champion, J. E., Rakow, A., Reeslund, K. L.,… Cole, D. A. (2009). Randomized controlled trial of a family cognitive–behavioral preventive intervention for children of depressed parents. Journal of Consulting and Clinical Psychology, 77, 1007–1020. Compas, B. E., Howell, D. C., Phares, V., Williams, R. A., & Giunta, C. T. (1989). Risk factors for emotional/behavioral problems in young adolescents: A prospective analysis of adolescent and parental stress and symptoms. Journal of Consulting and Clinical Psychology, 57, 732- 740. doi: 10.1037/0022–006X.57.6.732 Compas, B. E., Jaser, S. S., Dunbar, J. P., Watson, K. H., Bettis, A. H., Gruhn, M. A., & Willians, E. K. (2014). Coping and emotion regulation from childhood to early adulthood: Points of convergence and divergence. Australian Journal of Psychology, 66, 71–81 Compas, B. E., Orosan, P. G., & Grant, K. E. (1993). Adolescent stress and coping: Implications for psychopathology during adolescence. Journal of Adolescence, 16, 331–349. doi: 10.1006/jado.1993.1028 Compas, B. E., Worsham, N. L., & Ey, S. (1991). Conceptual and developmental issues in children's coping with stress. In A. La Greca, L. Siegel, J. Wallander, & C. E. Walker (Eds.), Stress and Coping in Child Health (7–24). New York, NY: Guilford Press. Connell, J. P., & Wellborn, J. G. (1990). Competence autonomy and relatedness: A motivational analysis of self-system processes. In R. Gunnar & L. A. Sroufe (Eds.), Self processes in development: Minnesota Symposium on Child Psychology, Vol. 23 (pp. 43–77). Chicago, IL: University of Chicago Press. Connor-Smith, J. K., Compas, B. E., Wadsworth, M. E., Thomsen, A. H., & Saltzman, H. (2000). Responses to stress in adolescence: Measurement of coping and involuntary stress responses. Journal of Counseling and Clinical Psychology, 68, 976–992. doi: 10.1037/0022–006X.68.6.976 Connor-Smith, J. K., & Flachsbart, C. (2007). Relations between personality and coping: A meta-analysis. Journal of Personality and Social Psychology, 93, 1080–1107. Conrod, P. J., Castellanos-Ryan, N. & Strang, J. (2010). Brief, personality-targeted coping

skills interventions and survival as a non-drug use over a 2-year period during adolescence. Archives of General Psychiatry, 67, 85–93. Contreras, J. M., & Kerns, K. A. (2000). Emotion regulation processes: Explaining links between parent-child attachment and peer relationships. In K. A. Kerns, J. M. Contreras& A. M. Neal-Barnett (Eds.), Family and peers: Linking two social worlds (pp. 1—25). Westport, CT: Praeger. Contreras, J. M., Kerns, K. A., Weimer, B. L., Gentzler, A. L., & Tomich, P. L. (2000). Emotion regulation as a mediator of associations between mother-child attachment and peer relationships in middle childhood. Journal of Family Psychology, 14, 111–124. Conway, C. C., Hammen, C., & Brennan, P. A. (2012). Expanding stress generation theory: Test of a transdiagnostic model. Journal of Abnormal Psychology, 121, 754–766. Coping Consortium (I. Sandler, B. Compas, T. Ayers, N. Eisenberg, E. A. Skinner, & P. Tolan) (Organizers) (1998, 2001). New conceptualizations of coping. Workshop sponsored by the Arizona State University Prevention Research Center. Tempe, AZ. Coyne, J. C., & Racioppo, M. W. (2000). Never the Twain shall meet? Closing the gap between coping research and clinical intervention research. American Psychologist, 55, 655– 664. doi: 10.1037/0003–066X.55.6.655. Creasey, G., & Hesson-McInnes, M. (2001). Affective responses, cognitive appraisals, and conflict tactics in adolecent romantic relationships: Associations with attachment orientations. Journal of Counseling Psychology, 48, 85–96. Crugnola, C. R., Tambelli, R., Spinelli, M., Gazzotti, S., Caprin, C., & Albizzati, A. (2011). Attachment patterns and emotion regulation strategies in the second year. Infant Behavior & Development, 34, 136–151. doi: 10.1016/j.infbeh.2010.11.002 Cummings, E. M., & Davies, P. T. (1999). Depressed parents and family functioning: Interpersonal effects and children's functioning and development. In T. Joiner T & J. C. Coyne (Eds.), The interactional nature of depression: Advances in interpersonal approaches (pp. 299–328). Washington, DC: American Psychological Association. Curtis, W. J., & Cicchetti, D. (2003). Moving research on resilience into the 21st century: Theoretical and methodological considerations in examining the biological contributors to resilience. Development and Psychopathology, 15, 773–810. Dagan, M., Sanderman, R., Hoff, C., Meijerink, W. J., Baas, P. C., van Haastert, M., & Hagedoorn, M. (2013, October). The interplay between partners' responsiveness and patients' need for emotional expression in couples coping with cancer. Journal of Behavioral Medicine, 1–11. doi: 10.1007/s10865–013–9543–4 Dahl, R. E. (2004). Adolescent brain development: a period of vulnerabilities and opportunities. Keynote address. Annals of the New York Academy of Sciences, 1021(1), 1–22.

Davis, D., Shaver, P. R., & Vernon, M. L. (2003). Physical, emotional, and behavioral reactions to breaking up: The roles of gender, age, emotional involvement, and attachment style. Personality and Social Psychology Bulletin, 29, 871–884. Dawson, G. Hessel, D., & Frey, K. (1994). Social influences n early-developing biological and behavioral systems related to risk for affective disorder. Developmental Psychopathology, 64, 759–779. Deci, E. L., & Ryan, R. M. (1985). Intrinsic motivation and self-determination in human behavior. New York, NY: Plenum Press. Decker, C. L. (2006). Coping in adolescents with cancer: A review of the literature. Journal of Psychosocial Oncology, 24, 123–140. doi: 10.1300/J077v24n04_07 DeLoache, J. S., Sugarman, S., & Brown, A. L. (1985). The development of error correction strategies in young children's manipulative play. Child Development, 3, 928–939. Denham, S. A. (1998). Emotional development in young children. New York, NY: Guilford Press. Dennis, T. A. (2010). Neurophysiological markers for emotion regulation from the perspective of emotion-cognition integration: Current directions and future challenges. Developmental Neuropsychology, 35, 212–230. DeOliveira, C. A., Bailey, H. N., Moran, G., & Pederson, D. R. (2004). Emotion socialization as a framework for understanding the development of disorganized attachment. Social Development, 13, 437–467. Derryberry, D., Reed, M. A., & Pilkenton-Taylor, C. (2003). Temperament and coping: Advantages of an individual differences perspective. Development and Psychopathology, 15, 1049–1066. doi: 10.1017/S0954579403000439 Derryberry, D., & Tucker, D. M. (2006). Motivation, self-regulation, and self-organization. In D. Cicchetti & D. J. Cohen (Eds), Developmental psychopathology: Developmental neuroscience (2nd ed.) (Vol. 2, pp. 502–532). Hoboken, NJ: Wiley. Diamond, A. (2002). Normal development of prefrontal cortex from birth to young adulthood: Cognitive functions, anatomy, and biochemistry. In D. Stuss & R. Knight (Eds.), Principles of frontal lobe function (pp. 466–503). New York, NY: Oxford University Press. doi: 10.1093/acprof:oso/9780195134971.003.0029 Diamond, A. (2013). Executive functions. Annual Review of Psychology, 64, 135–138. Diamond, A., Prevor, M. B., Callender, G., & Druin, D. P. (1997). Prefrontal cortex cognitive deficits in children treated early and continuously for PKU. Monographs of the Society for Research in Child Development, 62, 1–206 Diamond, L. M., & Aspinwall, L. G. (2003). Emotion regulation across the life span: An

integrative perspective emphasizing self-regulation, positive affect, and dyadic processes. Motivation and Emotions, 27, 125–156. Dodge, K. A., Greenberg, M. T., Malone, P. S., & the Conduct Problems Prevention Research Group. (2008). Testing an idealized dynamic cascade model of the development of serious violence in adolescence. Child Development, 79, 1907–1927. Donaldson, D., Prinstein, M. J., Danovsky, M., & Spirito, A. (2000). Patterns of children's coping with life stress: Implications for clinicians. American Journal of Orthopsychiatry, 70, 351–359. doi: 10.1037/h0087689 Dunn, J., Bretherton, I. & Munn, P. (1987). Conversations about feeling states between mothers and their young children. Developmental Psychology, 23, 132–139. Durston, S., Davidson, M. C., Tottenham, N., Galvan, A., Spicer, J., Fossella, J. A., & Casey, B. J. (2006). A shift from diffuse to focal cortical activity with development. Developmental Science, 9, 1–8. doi: 10.1111/j.1467–7687.2005.00454.x Dusek, J. B., & Danko, M. (1994). Adolescent coping styles and perceptions of parental child rearing. Journal of Adolescent Research, 9, 412–426. Dweck, C. S. (2008). Mindset: The new psychology of success. New York, NY: Random House. Ebata, A. T., & Moos, R. H. (1991). Coping and adjustment in distressed and healthy adolescents. Journal of Applied Developmental Psychology, 12, 33–54. doi: 10.1016/0193– 3973(91)90029–4 Egeland, B. (2007). Understanding developmental processes of resilience and psychopathology. In A. S. Masten (Ed.), Multilevel dynamics in developmental psychopathology: Pathways to the future: The Minnesota Symposia on Child Psychology (Vol. 34, pp. 83–118). Mahwah, NJ: Lawrence Erlbaum. Ein-Dor, T., Mikulincer, M., & Shaver, P. R. (2011). Attachment insecurities and the processing of threat-related information: Studying the schemas involved in insecure people's coping strategies. Journal of Personality and Social Psychology, 101, 78–93. Eisenberg, N., Fabes, R. A., Carlo, G., & Karbon, M. (1992). Emotional responsivity to others: Behavioral correlates and socialization antecedents. In N. Eisenberg & R. A. Fabes (Eds.), Emotion and its regulation in early development: New directions in child development (pp. 57–74). San Francisco, CA: Jossey-Bass. Eisenberg, N., Fabes, R. A., & Guthrie, I. K. (1997). Coping with stress: The roles of regulation and development. In S. A. Wolchik & I. N. Sandler (Eds.), Handbook of children's coping: Linking theory and intervention (pp. 41–70). New York, NY: Plenum Press. Eisenberg, N., Fabes, R. A., & Murphy, B. C. (1996). Parents' reactions to children's negative

emotions: Relations to children's social competence and comforting behavior. Child Development, 67, 2227–2247. Eisenberg, N., Fabes, R. A., Shepard, S. A., Murphy, B. C., Guthrie, I. K., Jones, S.,… & Maszk, P. (1997). Contemporaneous and longitudinal prediction of children's social functioning from regulation and emotionality. Child Development, 68, 642–664. doi: 10.1111/j.1467–8624.1997.tb04227.x Eisenberg, N., Valiente, C., & Sulik, M. J. (2009). How the study of regulation can inform the study of coping. New Directions in Child and Adolescent Development, 2009, 124, 75–86. doi: 10.1002/cd.241 Elias, M. J., Rothbaum, P. G., & Gara, M. (1986). Social-cognitive problem solving in children: Assessing the knowledge and application of skills. Journal of Applied Developmental Psychology, 7, 77–94. doi: 10.1016/0193–3973(86)90020–1 Ellis, B. J., Boyce, W. T., Belsky, J., Bakermans-Kranenburg, M. J., & Van IJzendoorn, M. H. (2011). Differential susceptibility to the environment: An evolutionary-neurodevelopmental theory. Development and Psychopathology, 23, 7. Essau, C. A., Conradt, J., Sasgawa, S., & Ollendick, T. H. (2012). Prevention of anxiety symptoms in children: Results from a universal school-based trial. Behavior Therapy, 43, 450–464. Evans, G. W. (2004). The environment of childhood poverty. American Psychologist, 59, 77– 92. Evans, G. W., & Kim, P. (2012). Childhood poverty and young adults' allostatic load: The mediating role of childhood cumulative risk exposure. Psychological Science, 23, 979–983. Eyberg, S. M., Schumann, E. M., & Rey, J. (1998). Child and adolescent psychology research: Developmental issues. Journal of Abnormal Child Psychology, 26, 71–82. Fabes, R. A., Eisenberg, N., & Bernzweig, J. (1990). The Coping with Children's Negative Emotions Scale: Description and scoring. Unpublished scale, Department of Family Resources and Human Development, Arizona State University. Fair, D. A., Cohen, A. L., Dosenbach, N. U., Church, J. A., Miezin, F. M., Barch, D. M.,… Schlaggar, B. L. (2008). The maturing architecture of the brain's default network. Proceedings of the National Academy of Sciences, 105(10), 4028–4032. doi: 10.1073/pnas.0800376105 Fan, J., Fossella, J., Sommer, T., Wu, Y., & Posner, M. I. (2003). Mapping the genetic variation of executive attention onto brain activity. Proceedings of the National Academy of Sciences of the United States of America, 100, 7406–7411. Farrell, L. J., Waters, A. M., & Zimmer-Gembeck, M. J. (2012). Cognitive biases and obsessive-compulsive symptoms in children: Examining the role of parent cognitive bias and

child age. Behavior Therapy, 43, 593–605. Fear, J. M., Champion, J. E., Reeslund, K. L., Forehand, R., Colletti, C., Roberts, L., & Compas, B. E. (2009). Parental depression and interparental conflict: Children and adolescents' self-blame and coping responses. Journal of Family Psychology, 23, 762. doi: 10.1037/a0016381 Feldman, R. (2009). The development of regulatory functions from birth to 5 years: Insights from premature infants. Child Development, 80, 544–561. doi: 10.1111/j.1467– 8624.2009.01278.x. Fields, L., & Prinz, R. J. (1997). Coping and adjustment during childhood and adolescence. Clinical Psychology Review, 17, 937–976. doi: 10.1016/S0272–7358(97)00033–0 Fischhoff, B. (2005). Development of an in behavioral decision research. In J. E. Jacobs & P. Klaczynski (Eds.), The development of judgment and decision-making in children and adults (pp. 335–346). Mahwah, NJ: Erlbaum. Folkman, S. (1984). Personal control and stress and coping processes: A theoretical analysis. Journal of Personality and Social Psychology, 46, 839–852. doi: 10.1037/0022– 3514.46.4.839 Folkman, S., & Lazarus, R. S. (1980). An analysis of coping in a middle-aged community sample. Journal of Health and Social Behavior, 21, 219–239. Folkman, S., & Lazarus, R. S. (1985). If it changes it must be a process: Study of emotion and coping during three stages of a college examination. Journal of Personality and Social Psychology, 48, 150–170. Folkman, S., & Moskowitz, J. T. (2004). Coping: Pitfalls and promise. Annual Review of Psychology, 55, 745–774. doi: 10.1146/annurev.psych.55.090902.141456 Fonagy, P., Gergely, G., & Target, M. (2007). The parent-infant dyad and the construction of the subjective self. Journal of Child Psychology and Psychiatry, 48, 228–328. Forsythe, C. J., & Compas, B. E. (1987). Interaction of cognitive appraisals of stressful events and coping: Testing the goodness of fit hypothesis. Cognitive Therapy and Research, 11, 473– 485. Fox, J. K., Masia Warner, C., Lerner, A. B., Ludwig, K., Ryan, J. L., Colognori, D.,… & Brotman, L. M. (2012). Preventive intervention for anxious preschoolers and their parents: Strengthening early emotional development. Child Psychiatry and Human Development, 43, 544–559. Fox, N. A., & Calkins, S. D. (2003). The development of self-control of emotion: Intrinsic and extrinsic influences. Motivation and emotion, 27, 7–26. doi: 10.1023/A:1023622324898 Fox, N. A., Henderson, H. A., Marshall, P. J., Nichols, K. E., & Ghera, M. M. (2005).

Behavioral inhibition: Linking biology and behavior within a developmental framework. Annual Review of Psychology, 56, 235–262. Fowler, P. J., Tompsett, C. J., Braciszewski, J. M., Jacques-Tiura, A. J., & Baltes, B. B. (2009). Community violence: A meta-analysis on the effect of exposure and mental health outcomes of children and adolescents. Development and Psychopathology, 21, 227–259. doi: 10.1017/S0954579409000145 Fraley, R. C., Waller, N. G., & Brennan, K. A. (2000). An item-response theory analysis of self-report measures of adult attachment. Journal of Personality and Social Psychology, 78, 350–365. Fresco, D. M., Mennin, D. S., Heimberg, R. G., & Ritter, M. (2013). Emotion regulation therapy for generalized anxiety disorder. Cognitive and Behavioral Practice, 20, 282–300. Freund, A. M. & Baltes, P. B. (1998). Selection, optimization, and compensation as strategies of life-management: Correlations with subjective indicators of successful aging. Psychology and Aging, 13, 531–543. Fridja, N. H. (1987). Emotions, cognitive structure, and action tendency. Cognition and Emotion, 1, 115–144. doi: 10.1080/02699938708408043 Frydenberg, E. (1997). Adolescent coping: Theoretical and research perspectives. New York, NY: Routledge. Frydenberg, E., & Lewis, R. (2000). Teaching coping to adolescents: When and to whom? American Educational Research Journal, 37, 727–745. Fuertes, M., Dos Santos, P. L., Beeghly, M., & Tronick, E. (2006). More than maternal sensitivity shapes attachment: Infant coping and temperament. Annals of the New York Academy of Science, 1094, 292–296. Galatzer-Levy, I. R., Burton, C. L., & Bonanno, G. A. (2012). Coping flexibility, potentially traumatic life events, and resilience: A prospective study of college student adjustment. Journal of Social and Clinical Psychology, 31, 542–567. doi: 10.1521/jscp.2012.31.6.542 Garber, J., Braafladt, N., & Weiss, B. (1995). Affect regulation in depressed and nondepressed children and young adolescents. Development and Psychopathology, 7, 93–115. doi: 10.1017/S0954579400006362 Garber, J., Carke, G. N., Weersing, V. R., Beardslee, W. R., Brent, D. A., Gladstone, T. R. G., … Iyengar, S. (2009). Prevention of depression in at-risk adolescents. Journal of the American Medical Association, 301, 2215–2224. Garmezy, N., & Rutter, M. (Eds.). (1988). Stress, coping and development in children. New York, NY: McGraw-Hill. Gaylord-Harden, N. K., Cunningham, J. A., Holmbeck, G. N., & Grant, K. E. (2010).

Suppressor effects in coping research with African American adolescents from low-income communities. Journal of Consulting and Clinical Psychology, 78, 843–855. Gaylord-Harden, N. K., Taylor, J. T., Campbell, C. L., Kesselring, C. M., & Grant, K. E. (2009). Maternal attachment and depressive symptoms in urban adolescents: The influence of coping strategies and gender. Journal of Clinical Child & Adolescent Psychology, 38, 684– 695. Gazelle, H. (2010). Anxious solitude/withdrawal and anxiety disorders: Conceptualization, co-occurrence, and peer processes leading toward and away from disorder in childhood. New Directions for Child and Adolescent Development, 127, 67–78. doi: 10.1002/cd.263 Gerardi, G., Rothbart, M. K., Posner, M. I., & Kelper. S. (1996). The development of attentional control: Performance on a spatial Stroop-like task at 24, 30 and 36–38 months of age. Poster session presented at the annual meeting of the International Society for Infant Studies. Providence, RI. Gilliom, M., Shaw, D. S., Beck, J. E., Schonberg, M. A., & Lukon, J. L. (2002). Anger regulation in disadvantaged preschool boys: Strategies, antecedents, and the development of self-control. Developmental Psychology, 38, 222–235. Ginsburg, G. S. (2009). The child anxiety prevention study: Intervention model and primary outcomes. Journal of Consulting and Clinical Psychology, 77, 580–587. Gogtay, N., Giedd, J. N., Lusk, L., Hayashi, K. M., Greenstein, D., Vaitiuzis, C.,… Thompson, P. M. (2004). Dynamic mapping of human cortical development during childhood and adolescence. Proceedings of the National Academy of Sciences, 101, 8174–8179. doi: 10.1073/pnas.0402680101 Gonzales, N. A., Dumka, L. E., Millsap, R. E., Gottschall, A., McClain, D. B., Wong, J. J.,… Kim, S. Y. (2012). Randomized trial of a broad preventive intervention for Mexican American adolescents. Journal of Consulting and Clinical Psychology, 80, 1–16. Gottman, J., Katz, L. F., & Hooven, C. (1996). Parent meta-emotion philosophy and the emotional life of families: Theoretical models and preliminary data. Journal of Family Psychology, 10, 243–268. Gottman, J., Katz, L. F., & Hooven, C. (1997). Meta-emotion: How families communicate emotionally. Mahwah, NJ: Erlbaum. Gottman, J. M., Katz, L. F., & Hooven, C. (2013). Meta-emotion: How families communicate emotionally. Mahwah, NJ: Lawrence Erlbaum. Gould, M. S., Velting, D., Kleinman, M., Lucas, C., Thomas, J. G., & Chung, M. (2004). Teenagers' attitudes about coping strategies and help-seeking behavior for suicidality. Journal of the American Academy of Child & Adolescent Psychiatry, 43, 1124–1133. doi: 10.1097/01.chi.0000132811.06547.31

Grant, K. E., Compas, B. E., Thurm, A. E., McMahon, S. D., Gipson, P. Y., Campbell, A. J.,… Westerholm, R. I. (2006). Stressors and child and adolescent psychopathology: Evidence of moderating and mediating effects. Clinical Psychology Review, 26, 257–283. Grimm, S., Schmidt, C. F., Bermpohl, F., Heinzel, A., Dahlem, Y., Wyss, M.,… Northoff, G. (2006) Segregated neural representation of distinct emotion dimensions in the prefrontal cortex —an fMRI study. NeuroImage, 30, 325–340. doi: 10.1016/j.neuroimage.2005.09.006 Grolnick, W. S. (2002). The psychology of parental control: How well-meant parenting backfires. Hillsdale, NJ: Erlbaum. Gross, J. J. (1998). Antecedent-and response-focused emotion regulation: Divergent consequences for experience, expression, and physiology. Journal of Personality and Social Psychology, 74(1), 224–237. doi: 10.1037/0022–3514.74.1.224 Gross, J. J., & Thompson, R. A. (2007). Emotion regulation: Conceptual foundations. In J. J. Gross (Ed.), Handbook of emotion regulation (pp. 3–24). New York, NY: Guilford Press. Grych, J. H., Fincham, F. D., Jouriles, E. N., & McDonald, R. (2000). Interparental conflict and child adjustment: testing the mediational role of appraisals in the cognitive-contextual framework. Child Development, 71, 1648–1661. doi: 10.1111/1467–8624.00255 Gunnar, M. R., & Cheatham, C. L. (2003). Brain and behavior interface: Stress and the developing brain. Infant Mental Health Journal, 24, 195–211. doi: 10.1002/imhj.10052 Gunnar, M. R., & Donzella, B. (2002). Social regulation of the cortisol levels in early human development. Psychoneuroendocrinology, 27, 199–220. Gunnar, M. R., Larson, M. C., Hertsgaard, L., Harris, M. L., & Brodersen, L. (1992). The stressfulness of separation among nine-month-old infants: Effects of social context variables and infant temperament. Child Development, 63, 290–303. doi: 10.1111/j.1467– 8624.1992.tb01627.x Gunnar, M. R., Marvinney, D., Isensee, J., & Fisch, R. O. (1989). Coping with uncertainty: New models of the relations between hormonal, behavioral, and cognitive processes. In D. S. Palermo, (Ed.), Coping with uncertainty: Behavioral and developmental perspectives (pp. 101–130). Hillsdale, NJ: Erlbaum. Gunnar, M. R., & Quevedo, K. (2007). The neurobiology of stress and development. Annual Review of Psychology, 58, 11.1–11.29. doi: 10.1146/annurev.psych.58.110405.085605 Gunnar, M. R., & Vazquez, D. (2006). Stress neurobiology and development psychopathology. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology: Developmental neuroscience (pp. 533–577) Hoboken, NJ: Wiley. Hagger, M. S., Wood, C., & Stiff, C. (2010). Ego depletion and the strength model of selfcontrol: A meta-analysis. Psychological Bulletin, 136, 495–525.

Hammen, C. (1999). The emergence of an interpersonal approach to depression. In T. Joiner and J. C. Coyne (Eds), The interactional nature of depression: Advances in interpersonal approaches, (pp. 21–35). Washington, DC: American Psychological Association. doi: 10.1037/10311–001 Hammen, C. (2005). Stress and depression. Annual Review of Clinical Psychology, 1, 293– 319. doi: 10.1146/annurev.clinpsy.1.102803.143938 Harman, C., Rothbart, M. K., & Posner, M. I. (1997). Distress and attention interactions in early infancy. Motivation and Emotion, 21, 27–44. Harper, B. D. (2012). Parents' and children's beliefs about peer victimization attributions, coping responses, and child adjustment. Journal of Early Adolescence, 32, 387–413. doi: 10.1177/0272431610396089 Harter, S. (2012). The construction of the self: Developmental and sociocultural foundations (2nd ed.). New York, NY: Guilford Press. Herbert, J., Goodyer, I. M., Grossman, A. B., Hastings, M. H., de Kloet, E. R., Lightman, S. L., & Seckl, J. R. (2006). Do corticosteroids damage the brain? Journal of Neuroendocrinology, 18, 393–411. doi: 10.1111/j.1365–2826.2006.01429.x Hoffman, M. L. (1994). Discipline and internalization. Developmental Psychology, 30, 26–28. doi: 1–.1037/0012-1649.30-1.25 Holahan, C. J., & Moos, R. H. (1991). Life stressors, personal and social resources, and depression: A 4-year structural model. Journal of Abnormal Psychology, 100, 31–38. doi: 10.1037/0021-843X.100.1.31 Holahan, C. J., Moos, R. H., Holahan, C. K., Brennan, P. L., & Schutte, K. K. (2005). Stress generation, avoidance coping, and depressive symptoms: A 10-year model. Journal of Consulting and Clinical Psychology, 73, 658- 666. doi: 10.1037/0022-006X.73.4.658 Holmberg, D., Lomore, C. D., Takacs, T. A., & Price, E. L. (2010). Adult attachment styles and stressor severity as moderators of the coping sequence. Personal Relationships, 18, 502–517. Holodynski, M., & Friedlmeier, W. (2006). Development of emotions and emotion regulation. New York, NY: Springer. Hornik, R., Risenhoover, N., & Gunnar, M. (1987). The effects of maternal positive, neutral, and negative affective communications on infant responses to new toys. Child Development, 58, 937–944. Horwitz, A. G., Hill, R. M., & King, C. A. (2011). Specific coping behaviors in relation to adolescent depression and suicidal ideation. Journal of Adolescence, 34, 1077–1085. doi: 10.1016/j.adolescence.2010.10.004 Izard, C. E., Hembree, E., & Huebner, R. (1987). Infants' emotional expressions to acute pain:

Developmental changes and stability of individual differences. Developmental Psychology, 23, 105–113. Jacobsen, P. B., Sadler, I. J., Booth-Jones, M., Soety, E., Weitzner, M. A., & Fields, K. K. (2002). Predictors of posttraumatic stress disorder symptomatology following bone marrow transplantation for cancer. Journal of Consulting and Clinical Psychology, 70, 235–240. doi: 10.1037/0022-006X.70.1.235 Jaffee, S. R., Caspi, A., Moffitt, T. E., Polo-Tomas, M., & Taylor, A. (2007). Individual, family, and neighbourhood factors distinguish resilient from non-resilient maltreated children: A cumulative stressors model. Child Abuse & Neglect, 31, 231–253. doi: 10.1016/j.chiabu.2006.03.011 Jaser, S. S., Champion, J. E., Reeslund, K. L., Keller, G., Merchant, M. J., Benson, M., & Compas, B. E. (2007). Cross-situational coping with peer and family stressors in adolescent offspring of depressed parents. Journal of Adolescence, 30(6), 917–932. doi: 10.1016/j.adolescence.2006.11.010 John, O. P., & Gross, J. J. (2004). Healthy and unhealthy emotion regulation: Personality processes, individual differences, and life span development. Journal of Personality, 72, 1301–1334. doi: 10.1111/j.1467-6494.2004.00298.x Jones, L. B., Rothbart, M. K., & Posner, M. I. (2003). Development of executive attention in preschool children. Developmental Science, 6, 498–504. doi: 10.1111/1467-7687.00307 Jones, S., Eisenberg, N., Fabes, R. A., & MacKinnon, D. P. (2002). Parents' reactions to elementary school children's negative emotions: Relations to social and emotional functioning at school. Merrill-Palmer Quarterly, 48, 133–159. Kagan, J. (1997). Temperament and the reactions to unfamiliarity. Child Development, 68, 139–143. Kaminsky, L., Robertson, M., & Dewey, D. (2006). Psychological correlates of depression in children with recurrent abdominal pain. Journal of Pediatric Psychology, 31, 956–966. doi: 10.1093/jpepsy/jsj103 Kazdin, A. E. (2008). Evidence-based treatment and practice: New opportunities to bridge clinical research and practice, enhance the knowledge base, and improve patient care. American Psychologist, 63, 146–159. doi: 10.1037/0003-066X.63.3.146 Keen, R. (2011). The development of problem solving in young children: A critical cognitive skill. Annual Review of Psychology, 62, 1–21. Keenan, K. (2000). Emotion dysregulation as a risk factor for psychopathology. Clinical Psychology: Science and Practice, 7, 418–434. Kendall, E., & Terry, D. J. (2008). Understanding adjustment following traumatic brain injury:

Is the Goodness-of-Fit coping hypothesis useful? Social Science & Medicine, 67, 1217–1224. doi: 10.1016/j.socscimed.2008.05.033 Khurana, A., & Romer, D. (2012). Modeling the distinct pathways of influence of coping strategies on youth suicidal ideation: A national longitudinal study. Prevention Science, 13, 644–654. doi: 10.1007/s11121-012-0292-3 Kingsbury, M., Coplan, R. J., & Rose-Krasnor, L. (2013). Shy but getting by? An examination of the complex links among shyness, coping, and socioemotional functioning in childhood. Social Development, 22, 126–145. doi: 10.1111/sode.12003 Kliewer, W., Fearnow, M. D., & Miller, P. A. (1996). Coping socialization in middle childhood: Tests of maternal and paternal influences. Child Development, 67, 2339–2357. Kliewer, W., Parrish, K. A., Taylor, K. W., Jackson, K., Walker, J. M., & Shivy, V. A. (2006). Socialization of coping with community violence: Influences of caregiver coaching, modeling, and family context. Child Development, 77, 605–623. Kliewer, W., Sandler, I., & Wolchik, S. (1994). Family socialization of threat appraisal and coping: Coaching, modeling, and family context. In K. Hurrelman & F. Nestmann (Eds.), Social networks and social support in childhood and adolescence (pp. 271–291). Berlin, Germany: de Gruyter. Kobak R., Cassidy, J., Lyons-Ruth, K. & Zir, Y. (2006). Attachment, stress and psychopathology: A developmental pathways model. In D. Cicchetti, and Cohen (Eds.), Handbook of developmental psychopathology, Vol. 1 (pp. 333—369). New York, NY: Cambridge University Press. Kobak, R., Little, M., Race, E., & Acosta, M. (2001). Attachment disruptions in seriously emotionally disturbed children: Implications for treatment. Attachment and Human Development, 3, 243–258. Kochanska, G., Coy, K. C., & Murray, K. T. (2001). The development of self-regulation in the first four years of life. Child Development, 72, 1091–1111. Kochanska, G., Forman, D. R., Aksan, N., & Dunbar, S. B. (2005). Pathways to conscience: Early mother-child mutually responsive orientation and children's moral emotion, conduct, and cognition. Journal of Child Psychology and Psychiatry, 46, 19–34. Kochanska, G., & Kim, S. (2013). A complex interplay among the parent–child relationship, effortful control, and internalized, rule-compatible conduct in young children: Evidence from two studies. Child Development, 84, 283–296. doi: 10.1111/j.1467-8624.2012.01852.x Kochanska, G., Philibert, R. A., & Barry, R. A. (2009). Interplay of genes and early mother– child relationship in the development of self-regulation from toddler to preschool age. Journal of Child Psychology and Psychiatry, 50, 1331–1338. doi: 10.1111/j.1469-7610.2008.02050.x

Kochenderfer-Ladd, B., & Skinner, K. (2002). Children's coping strategies: Moderators of the effects of peer victimization? Developmental Psychology, 38, 267. doi: 10.1037/00121649.38.2.267 Kopp, C. B. (1989). Regulation of distress and negative emotions: A developmental view. Developmental Psychology, 25, 343–354. Kopp, C. B. (2002). Commentary: The co-developments of attention and emotion regulation. Infancy, 2, 199–208. Krattenmacher, T., Kühne, F., Führer, D., Beierlein, V., Brähler, E., Resch, F.,… & Möller, B. (2013). Coping skills and mental health status in adolescents when a parent has cancer: A multicenter and multi-perspective study. Journal of Psychosomatic Research, 74, 252–259. doi: 10.1016/j.jpsychores.2012.10.003 Kremen, A. M., & Block, J. (1998). The roots of ego-control in young adulthood: Links with parenting in early childhood. Journal of Personality and Social Psychology, 75, 1062–1075. Kroger, J. (2007). Identity development: Adolescence through adulthood. Newbury Park, CA: Sage. Kuhn, D., & Franklin, S. (2006). The second decade: What develops (and how)? In W. Damon & R. Lerner (Series Eds.), D. Kuhn & R. Siegler (Vol. Eds.), Wiley Handbook of child psychology: Vol. 2. Cognition, perception, and language (6th ed., pp. 962–993). Hoboken, NJ: Wiley Kullik, A., & Petermann, F. (2013). Attachment to parents and peers as a risk factor for adolescent depressive disorders: The mediating role of emotion regulation. Child Psychiatry and Human Development, 44, 537–548. doi: 10.1007/s10578-012-0347-5 Laible, D. J., & Thompson, R. A. (1998). Attachment and emotional understanding in preschool children. Developmental Psychology, 34, 1038–1045 Langner, T. S. (1962). A twenty-two item screening score of psychiatric symptoms indicating impairment. Journal of Health and Human Behavior, 3, 269–276. Langrock, A. M., Compas, B. E., Keller, G., Merchant, M. J., & Copeland, M. E. (2002). Coping with the stress of parental depression: Parents' reports of children's coping, emotional, and behavioral problems. Journal of Clinical Child and Adolescent Psychology, 31, 312– 324. doi: 10.1207/S15374424JCCP3103_03 Lansford, J. E., Malone, P. S., Dodge, K. A., Pettit, G. S., & Bates, J. E. (2010). Developmental cascades of peer rejection, social information processing biases, and aggression during middle childhood. Development and Psychopathology, 22, 593–602. Larson, R. W. (2011). Adolescents' conscious processes of developing regulation: Learning to appraise challenges. In R. M. Lerner, J. V. Lerner, E. P. Bowers, S. Lewin-Bizan, S.

Gestsdottir, & J. B. Urban (Eds.), Thriving in childhood and adolescence: The role of selfregulation processes. New Directions for Child and Adolescent Development, 133, 87–97. Lazarus, R. S. (1994). Emotion and adaptation. New York, NY: Oxford University Press. Lazarus, R. S. (2000). Toward better research on stress and coping. American Psychologist, 55(6), 665–673. doi: 10.1037/0003-066X.55.6.665 Lazarus, R. S., & Folkman, S. (1984). Stress, appraisal, and coping. New York, NY: Springer. Lazarus, R. S., & Folkman, S. (1986). Cognitive theories of stress and the issue of circularity. In M. H. Appley & R. Trumbull (Eds.), Dynamics of stress: Physiological, psychologcal, and social perspectives (pp. 63–80). New York, NY: Plenum Press. Lengua, L. J., & Long, A. C. (2002). The role of emotionality and self-regulation in the appraisal–coping process: Tests of direct and moderating effects. Journal of Applied Developmental Psychology, 23(4), 471–493. doi: 10.1016/S0193-3973(02)00129-6 Lengua, L. J., Sandler, I. N., West, S. G., Wolchik, S. A., & Curan, P. J. (1999). Emotionality and self-regulation, threat appraisal, and coping in children of divorce. Development and Psychopathology, 11, 15–37. Lengua, L. J., & Stormshak, E. A. (2000). Gender, gender roles, and personality: Gender differences in the prediction of coping and psychological symptoms. Sex Roles, 43, http://link.springer.com/journal/11199/43/11/page/1787–820. doi: 10.1023/A:1011096604861 Lewis, M. (1997). The self in self-conscious emotions. Annals of the New York Academy of Sciences, 818, 119–142. Lewis, M. D., & Todd, R. M. (2007). The self-regulating brain: Cortical-subcortical feedback and the development of intelligent action. Cognitive Development, 22, 406–430. doi: 10.1016/j.cogdev.2007.08.004 Li, C. E., DiGiuseppe, R., & Froh, J. (2006). The roles of sex, gender, and coping in adolescent depression. Adolescence, 41, 409–415. Littleton, H., Axsom, D., & Grills-Taquechel, A. E. (2011). Longitudinal evaluation of the relationship between maladaptive trauma coping and distress: Examination following the mass shooting at Virginia Tech. Anxiety, Stress, & Coping, 24, 273–290. doi: 10.1080/10615806.2010.500722 Liu, R. T. (2013). Stress generation: Future directions and clinical implications. Clinical Psychology Review, 33, 406–416. Loman, M. M., & Gunnar, M. R. (2010). Early experience and the development of stress reactivity and regulation in children. Neuroscience & Biobehavioral Reviews, 34, 867–876. doi: 10.1016/j.neubiorev.2009.05.007

Lopez, F. G., Mauricio, A. M., Gormley, B., Simko, T., & Berger, E. (2001). Adult attachment orientations and college student distress: The mediating role of problem coping styles. Journal of Counseling & Development, 79, 459–464. Losoya, S., Eisenberg, N., & Fabes, R. A. (1998). Developmental issues in the study of coping. International Journal of Behavioral Development, 22, 287–313. doi: 10.1080/016502598384388 Luciana, M., & Nelson, C. A. (1998). The functional emergence of prefrontally-guided working memory systems in four-to eight-year-old children. Neuropsychologia, 36, 273–293. doi: 10.1016/S0028-3932(97)00109-7 Luna, B., Padmanabhan, A., & O'Hearn, K. (2010). What has fMRI told us about the development of cognitive control through adolescence?. Brain and Cognition, 72, 101–113. doi: 10.1016/j.bandc.2009.08.005 Luna, B., & Sweeney, J. A. (2004). The emergence of collaborative brain function: FMRI studies of the development of response inhibition. Annals of the New York Academy of Sciences, 1021, 296–309. doi: 10.1016/j.bandc.2009.08.005 Luna, B., Thulborn, K. R., Munoz, D. P., Merriam, E. P., Garver, K. E., Minshew, N. J.,… Sweeney, J. A. (2001). Maturation of widely distributed brain function subserves cognitive development. Neuroimage, 13, 786–793. doi: 10.1006/nimg.2000.0743 Luthar, S. S. (2006). Resilience in development: A synthesis of research across five decades. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Risk, disorder, and adaptation (2nd ed., Vol. 3, pp. 739–795). Hoboken, NJ: Wiley. Lynch, M., & Cicchetti, D. (1998). An ecological-transactional analysis of children and contexts: The longitudinal interplay among child maltreatment, community violence, and children's symptomatology. Development and Psychopathology, 10, 235–257. Malatesta, C. Z., Culver, C., Tesman, J. R., & Shepard, B. (1989). The development of emotion expression during the first two years of life. Monographs of the Society for Research in Child Development, 54, 1–104. Markovic, A., Rose-Krasnor, L., & Coplan, R. (2013). Children's coping with social conflict: The role of personality self-theories. Personality and Individual Differences, 54, 64–69. Marshall, P. (2013). Coping with complexity: Developmental systems and multilevel analyses in developmental psychopathology. Development and Psychopathology, 25, 1311–1324. Martins, E. C., Soares, I., Martins, C., Tereno, S., & Osorio, A. (2012). Can we identify emotion over-regulation in infancy? Associations with avoidant attachment, dyadic emotional interaction and temperament. Infant and Child Development, 21, 579–595. Masten, A. S. (2001). Ordinary magic: Resilience processes in development. American

Psychologist, 56, 227–238. doi: 10.1037/0003-066X.56.3.227 Masten, A. S. (2007). Resilience in developing systems: Progress and promise as the fourth wave rises. Development and Psychopathology, 19, 921–930. Masten, A. S., & Cicchetti, D. (2010). Developmental cascades. Development and Psychopathology, 22, 491–495. Mattlin, J. A., Wethington, E., & Kessler, R. C. (1990). Situational determinants of coping and coping effectiveness. Journal of Health and Social Behavior, 31, 103–122. doi: 10.2307/2137048 Maughan, A., & Cicchetti, D. (2002). Impact of child maltreatment and interadult violence on children's emotion regulation abilities and socioemotional adjustment. Child Development, 73, 1525–1542. McCarthy, C. J., Lambert, R. G., & Seraphine, A. (2004). Adaptive family functioning and emotion regulation capacities as predictors of college students' appraisals and emotion valence following conflict with their parents. Cognition and Emotion, 18, 97–124. McCarty, M. E., Clifton, R. K., & Collard, R. R. (1999). Problem solving in infancy: The emergence of an action plan. Developmental Psychology, 35, 1091–1101. McClure, E. B., & Pine, D. S. (2006). Social anxiety and emotion regulation: A model for developmental psychopathology perspectives on anxiety disorders. In D. Cicchetti and Cohen, D. J. (Eds.), Handbook of developmental psychopathology, Vol. 3 (pp. 470–502). New York, NY: Cambridge University Press. McEwen, B. S. (1998). Seminars in medicine of the Beth Israel Deaconess Medical Center: Protective and damaging effects of stress mediators. New England Journal of Medicine, 338, 171–179. McEwen, B. S. (2004). Protection and damage from acute and chronic stress: Allostasis and allostatic overload and relevance to the pathophysiology of psychiatric disorders. Annals of the New York Academy of Sciences, 1032, 1–7. doi: 10.1196/annals.1314.001 McEwen, B. S. (2010). Stress: Homeostasis, rheostasis, allostasis and allostatic load. Stress Science: Neuroendocrinology, 10–14. McLaughlin, K. A., & Nolen-Hoeksema, S. (2011). Rumination as a transdiagnostic factor in depression and anxiety. Behaviour Research and Therapy, 49, 186–193. Meeus, W. (2011). The study of adolescent identity formation 2000–2010: A review of longitudinal research. Journal of Research on Adolescence, 21, 75–94. Mesquita, B., & Albert, D. (2007). The cultural regulation of emotions. In. J. J. Gross (Ed.), Handbook of emotion regulation (pp. 486–503). New York, NY: Guilford Press.

Metcalfe, J., & Mischel, W. (1999). A hot/cool-system analysis of delay of gratification: Dynamics of willpower. Psychological Review, 106, 3–19. Mezulis, A., Hyde, J. S., & Abramson, L. Y. (2006). The developmental origins of cognitive vulnerability to depression: Temperament, parenting, and negative life events in childhood as contributors to negative cognitive style. Developmental Psychology, 42, 1012–1025. doi: 10.1037/0012-1649.42.6.1012 Mezulis, A., Simonson, J., McCauley, E., & Vander Stoep, A. (2011). The association between temperament and depressive symptoms in adolescence: Brooding and reflection as potential mediators. Cognition & Emotion, 25, 1460–1470. doi: 10.1080/02699931.2010.543642 Michl, L. C., McLaughlin, K. A., Shepherd, K., & Nolen-Hoeksema, S. (2013). Rumination as a mechanism linking stressful life events to symptoms of depression and anxiety: Longitudinal evidence in early adolescents and adults. Journal of Abnormal Psychology, 122, 339–352. doi: 10.1037/s0031994 Mikulincer, M., & Florian, V. (2003). Attachment style and affect regulation: Implications for coping with stress and mental health. In G. J. O. Fletcher & M. S. Clark (Eds.), Blackwell handbook of social psychology: Interpersonal processes (pp. 537–557). New York, NY: Blackwell. Miller, K. S., Vannatta, K., Compas, B. E., Vasey, M., McGoron, K. D., Salley, C. G., & Gerhardt, C. A. (2009). The role of coping and temperament in the adjustment of children with cancer. Journal of Pediatric Psychology, 34, 1135–1143. Miller, P. J., & Sperry, L. L. (1987). The socialization of anger and aggression. MerrillPalmer Quarterly, 33, 1–31. Miller, S. M. (1992). Individual differences in the coping process: What to know and when to know it. In B. N. Carpenter (Ed.), Personal coping: Theory, research, and application (pp. 77–91). Westport, CT: Praeger/Greenwood. Miller, S. M., & Green, M. L. (1985). Coping with stress and frustration. In M. Lewis & C. Saarni (Eds.), The socialization of emotions (pp. 263–314). New York, NY: Plenum Press. doi: 10.1007/978-1-4613-2421-8_12 Moos, R. H., & Holahan, C. J. (2003). Dispositional and contextual perspectives on coping: Toward an integrative framework. Journal of Clinical Psychology, 59, 1387–1403. doi: 10.1002/jclp.10229 Morales, M., & Bridges, L. J. (1996). Associations between nonparental care experience and preschooler's emotion regulation in the presence of the mother. Journal of Applied Developmental Psychology, 17, 577–596. Morris, A. S., Silk, J. S., Steinberg, L., Myers, S. S., & Robinson, L. R. (2007). The role of the family context in the development of emotion regulation. Social Development, 16, 361–388.

doi: 10.1111/j.1467-9507.2007.00389.x. Murphy, L., & Moriarity, A. (1976). Vulnerability, coping, and growth: From infancy to adolescence. New Haven, CT: Yale University Press. Nachmias, M., Gunnar, M., Mangelsdorf, S., Parritz, R. H., & Buss, K. (1996). Behavioral inhibition and stress reactivity: The moderating role of attachment security. Child Development, 67, 508–522. doi: 10.1111/j.1467-8624.1996.tb01748.x Neitzel, C., & Stright, A. D. (2003). Mothers' scaffolding of children's problem solving: establishing a foundation of academic self-regulatory competence. Journal of Family Psychology, 17, 147–159. Nes, L. S., & Segerstrom, S. C. (2006). Dispositional optimism and coping: A meta-analytic review. Personality and Social Psychology Review, 10, 235–251. doi: 10.1207/s15327957pspr1003_3 Newman, R. S., Murray, B., & Lussier, C. (2001). Confrontation with aggressive peers at school: Students' reluctance to seek help from the teacher. Journal of Educational Psychology, 93, 398–410. Nieder, T., & Seiffge-Krenke, I. (2001). Coping with stress in different phases of romantic development. Journal of Adolescence, 24, 297–311. doi: 10.1006/jado.2001.0407 Nigg, J. T. (2006). Temperament and developmental psychopathology. Journal of Child Psychology and Psychiatry, 47, 395–422. Nolen-Hoeksema, S., Girgus, J., & Seligman, M. E. P. (1992). Predictors and consequences of childhood depressive symptoms. Journal of Abnormal Psychology, 101, 405–422. Nolen-Hoeksema, S., Wisco, B., & Lyubomirsky, S. (2008). Rethinking rumination. Perspectives on Psychological Science, 3, 400–424. Nolte, T., Guiney, J., Fonagy, P., Mayes, L. C. & Luyten, P. (2011). Interpersonal stress regulation and the development of anxiety disorders: An attachment-based developmental framework. Frontiers in Behavioral Neuroscience, 5, 55–138. Ohman, A., & Mineka, S. (2001). Fears, phobias, and preparedness: Toward an evolved module of fear and fear learning. Psychological Review, 108, 483–522. O'Leary, A. (1990). Stress, emotion, and human immune function. Psychological Bulletin, 108, 363–382. doi: 10.1037/0033-2909.108.3.363. Panak, W. F., & Garber, J. (1992). Role of aggression, rejection, and attributions in the prediction of depression in children. Development and Psychopathology, 4, 145–165. doi: 10.1017/S0954579400005617 Park, C. L., Armeli, S., & Tennen, H. (2004). Appraisal-coping goodness of fit: A daily

Internet study. Personality and Social Psychology Bulletin, 30, 558–569. doi: 10.1177/0146167203262855 Parker, J. G., Rubin, K. H., Price, J. M., & DeRosier, M. E. (1995). Peer relationships, child development, and adjustment: A developmental psychopathology perspective. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, Vol. 2 (pp. 419–493) Hoboken, NJ: Wiley. Pascoe, E. A., & Smart Richman, L. (2009). Perceived discrimination and health: A metaanalytic review. Psychological Bulletin, 135, 531–554. doi: 10.1037/a0016059 Pellegrini, A. D., & Bartini, M. (2001). Dominance in early adolescent boys: Affiliative and aggressive dimensions and possible functions. Merrill-Palmer Quarterly, 47, 142.163. Penley, J. A., Tomaka, J., & Wiebe, J. S. (2002). The association of coping to physical and psychological health outcomes: A meta-analytic review. Journal of Behavioral Medicine, 25, 551–603. doi: 10.1023/A:1020641400589 Pennington, B. F., & Ozonoff, S. (1996). Execuive functions and developmental psychopathology. Journal of Child Psychology and Psychiatry, 37, 51–87. Petticrew, M., Bell, R., & Hunter, D. (2002). Influence of psychological coping on survival and recurrence in people with cancer: Systematic review. British Medical Journal, 325, 1066–1076. doi: 10.1136/bmj.325.7372.1066 Power, T. G. (2004). Stress and coping in childhood: The parents' role. Parenting: Science and Practice, 4, 271–317. Prencipe, A., & Zelazo, P. D. (2005). Development of affective decision making for self and other evidence for the integration of first- and third-person perspectives. Psychological Science, 16, 501–505. doi: 10.1111/j.0956-7976.2005.01564.x Prinstein, M. J., & La Greca, A. M. (2002). Peer crowd affiliation and internalizing distress in childhood and adolescence: A longitudinal follow-back study. Journal of Research on Adolescence, 12, 325–351. doi: 10.1111/1532-7795.00036 Raby, K. L., Cicchetti, D., Carlson, E. A., Cutuli, J. J., Englund, M. M., & Egeland, B. (2012). Genetic and caregiving-based contributions to infant attachment: Unique associations with distress reactivity and attachment security. Psychological Science, 23, 1016–1023. Rafnsson, F. D., Jonsson, F. H., & Windle, M. (2006). Coping strategies, stressful life events, problem behaviors, and depressed affect. Anxiety, Stress, and Coping, 19, 241–257. doi: 10.1080/10615800600679111 Rayburn, N. R., Wenzel, S. L., Elliott, M. N., Hambarsoomians, K., Marshall, G. N., & Tucker, J. S. (2005). Trauma, depression, coping, and mental health service seeking among impoverished women. Journal of Consulting and Clinical Psychology, 73, 667–677. doi:

10.1037/0022-006X.73.4.667 Repetti, R. L., Taylor, S. E., & Seeman, T. E. (2002). Risky families: Family social environments and the mental and physical health of offspring. Psychological Bulletin, 128(2), 330. Reyna, V. F., & Farley, F. (2006). Risk and rationality in adolescent decision-making: Implications for theory, practice, and public policy. Psychological Science in the Public Interest, 7, 1–44. doi: 10.1111/j.1529-1006.2006.00026.x Roemer, L., Orsillo, S. M., & Salters-Pedneault, K. (2008). Efficacy of an acceptance-based therapy for generalized anxiety disorders: Evaluation of a randomized controlled trial. Journal of Consulting and Clinical Psychology, 76, 1083–1089. Roesch, S. C., & Weiner, B. (2001). A meta-analytic review of coping with illness: Do causal attributions matter? Journal of Psychosomatic Research, 50, 205–219. doi: 10.1016/S00223999(01)00188-X Rohde, P., Lewinsohn, P. M., Tilson, M., & Seeley, J. R. (1990). Dimensionality of coping and its relation to depression. Journal of Personality and Social Psychology, 58, 499–511. doi: 10.1037/0022-3514.58.3.499 Roque, L., Verissimo, M., Fernandes, M., & Rebelo, A. (2013). Emotion regulation and attachment: Relationships with children's secure base during different situational and social contexts in naturalistic settings. Infant Behavior and Development, 36, 298–306. Rossman, B. B. R. (1992). School-aged children's perceptions of coping with distress: Strategies for emotion regulation and the moderation of adjustment. Journal of Child Psychiatry, 33, 1373–1397. Roth, S., & Cohen, L. (1986). Approach, avoidance, and coping with stress. American Psychologist, 41, 813–819. Rothbart, M. K. (2011). Becoming who we are: Temperament and personality in development. New York, NY: Guilford Press. Rothbart, M. K., Derryberry, D., & Posner, M. I. (1994). A psychobiological approach to the development of temperament. In J. E. Bates & T. D. Wachs (Eds.), Temperament: Individual differences at the interface of biology and behavior (pp. 83–116). Washington, DC: American Psychological Association. Rothbart, M. K., & Posner, M. I. (2006). Temperament, attention, and developmental psychopathology. Hoboken, NJ: Wiley. Rothbart, M. K., Posner, M. I., & Kieras, J. (2006). Temperament, attention, and the development of self-regulation. In K. McCartney & D. Phillips (Eds.), Blackwell handbook of early childhood development (pp. 338–357). Oxford, UK: Blackwell.

Rothbaum, F., Weisz, J. R., & Snyder, S. S. (1982). Changing the world and changing the self: A two-process model of perceived control. Journal of Personality and Social Psychology, 42, 5–37. Rubia, K., Overmeyer, S., Taylor, E., Brammer, M., Williams, S. C. R., Simmons, A., & Bullmore, E. T. (2000). Functional frontalisation with age: Mapping neurodevelopmental trajectories with fMRI. Neuroscience & Biobehavioral Reviews, 24, 13–19. doi: 10.1016/S0149-7634(99)00055-X Rudolph, K. D., & Asher, S. R. (2000). Adaptation and maladaptation in the peer system. In M. Lewis & S. M. Miller (Eds.), Handbook of developmental psychopathology (pp. 157–175). New York, NY: Springer. Rudolph, K. D., Dennig, M. D., & Weisz, J. R. (1995). Determinants and consequences of children's coping in the medical setting: Conceptualization, review, and critique. Psychological Bulletin, 118, 328–357. Rueda, M. R., & Rothbart, M. K. (2009). The influence of temperament on the development of coping: The role of maturation and experience. New Directions in Child and Adolescent Development, 2009, 19–32. doi: 10.1002/cd.239 Ruff, H. A., & Rothbart, M. K. (1996) Attention in early development: Themes and variations. New York, NY: Oxford University Press. Rutter, M. (2005). Multiple meanings of a developmental perspective on psychopathology. European Journal of Developmental Psychology, 2(3), 221–252. Rutter, M. (2013). Annual research review: Resilience–clinical implications. Journal of Child Psychology and Psychiatry, 54(4), 474–487. Ryan, R. M. (1992). Agency and organization: Intrinsic motivation, autonomy, and the self in psychological development. In J. Jacobs (Ed.), Nebraska symposium on motivation (Vol. 40, pp. 1–56). Lincoln, NE: University of Nebraska Press. Ryan, R. M., Deci, E. L., Grolnick, W. S., & La Guardia, J. G. (2006). The significance of autonomy and autonomy support in psychological development and psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Theory and method (2nd ed., Vol. 1, pp. 795–849). Hoboken, NJ: Wiley. Saarni, C. (1997). Coping with aversive feelings. Motivation and Emotion, 21, 45–63. doi: 10.1002/9780470147658.chpsy0305 Sameroff, A. (2010). A unified theory of development: A dialectic integration of nature and nurture. Child Development, 81, 6–22. Sameroff, A. J., & Haith, M. M. (1996). Interpreting developmental transitions. In A. J. Sameroff & M. M. Haith (Eds.), The five to seven year shift: The age of reason and

responsibility (pp. 3–15). Chicago, IL: University of Chicago Press. Sandler, I. N., Kim-Bae, L. S., & MacKinnon, D. (2000). Coping and negative appraisal as mediators between control beliefs and psychological symptoms in children of divorce. Journal of Clinical Child Psychology, 29, 336–347. doi: 10.1207/S15374424JCCP2903_5 Sandler, I. N., Tein, J.-Y., Mehta, P., Wolchik, S., & Ayers, T. (2000). Coping efficacy and psychological problems of children of divorce. Child Development, 71, 1099–1118. doi: 10.1111/1467-8624.00212 Sandler, I. N., Wolchik, S. A., MacKinnon, D., Ayers, T. S., & Roosa, M. W. (1997). Developing linkages between theory and intervention in stress and coping processes. In S. A Wolchik & I. N. Sandler (Eds.), Handbook of children's coping: Linking theory and intervention (pp. 3–40). New York, NY: Plenum Press. Sandstrom, M. J. (2004). Pitfalls of the peer world: How children cope with common rejection experiences. Journal of Abnormal Child Psychology, 32, 67–81. doi: 10.1023/B:JACP.0000007581.95080.8b Sansom-Daly, U. M., Peate, M., Wakefield, C. E., Bryant, R. A., & Cohn, R. J. (2012). A systematic review of psychological interventions for adolescents and young adults living with chronic illness. Health Psychology, 31, 380–393. Santiago-Rivera, A. L., Bernstein, B. L., & Gard, T. L. (1995). The importance of achievement and the appraisal of stressful events as predictors of coping. Journal of College Student Development, 36, 374–383. Sapolsky, R. M. (1992). Stress, the aging brain, and the mechanisms of neuron death. Cambridge, MA: MIT Press. Sapolsky, R. M. (1999). Glucocorticoids, stress, and their adverse neurological effects: Relevance to aging. Experimental Gerontology, 34, 721–732. doi: 10.1016/S05315565(99)00047-9. doi: 10.1016/S0531-5565(99)00047-9 Schwarz, B., Stutz, M., & Ledermann, T. (2012). Perceived interparental conflict and early adolescents' friendships: The role of attachment security and emotion regulation. Journal of Youth and Adolescence, 41, 1240–1252. doi: 10.1007/s10964-012-9769-4 Seeman, T., Epel, E., Gruenewald, T., Karlamangla, A., & McEwen, B. S. (2010). Socioeconomic differentials in peripheral biology: Cumulative allostatic load. Annals of the New York Academy of Sciences, 1186, 223–239. Seiffge-Krenke, I. (1995). Stress, coping and relationships in adolescence. Hillsdale, NJ: Erlbaum. Seiffge-Krenke, I. (2006). Coping with relationship stressors: The impact of different working models of attachment and links to adaptation. Journal of Youth and Adolescence, 35, 25–39.

Seiffge-Krenke, I. (2011). Coping with relationship stressors: A decade review. Journal of Research on Adolescence, 21, 196–210. doi: 10.1111/j.1532-7795.2010.00723.x Seiffge-Krenke, I., & Klessinger, N. (2000). Long-term effects of avoidant coping on adolescents' depressive symptoms. Journal of Youth and Adolescence, 29, 617–630. doi: 10.1023/A:1026440304695 Shah, B. M., Gupchup, G. V., Borrego, M. E., Raisch, D. W., & Knapp, K. K. (2012). Depressive symptoms in patients with Type 2 Diabetes Mellitus: Do stress and coping matter? Stress and Health, 28, 111–122. doi: 10.1002/smi.1410 Shannon, K. E., Beauchaine, T. P., Brenner, S. L., Neuhaus, E., & Gatzke-Kopp, L. (2007). Familial and temperamental predictors of resilience in children at risk for conduct disorder and depression. Development and Psychopathology, 19, 701. doi: 10.1017/S0954579407000351 Shaver, P. R., & Mikulincer, M. (2007). Adult attachment strategies and the regulation of emotion. In J. J. Gross (Ed.), Handbook of emotion regulation (pp. 446–465). New York, NY: Guilford Press. Shaw, D. S., Winslow, E. B., Owens, E. B., Vondra, J. I., Cohn, J. F., & Bell, R. Q. (1998). The development of early externalizing problems among children from low-income families: A transformational perspective. Journal of Abnormal Child Psychology, 26, 95–107. Shimazu, A., & Schaufeli, W. B. (2007). Does distraction facilitate problem-focused coping with job stress? A 1 year longitudinal study. Journal of Behavioral Medicine, 30, 423–434. doi: 10.1007/s10865-007-9109-4 Shipman, K. L., Schneider, R., Fitzgerald, M. M., Sims, C., Swisher, L, & Edwards, A (2007). Maternal emotion socialization in maltreating and nonmaltreating families: implications for children's emotion regulation. Social Development, 16, 268–285. Silverman, W. K., Kurtines, W. M., Jaccard, J., & Pina, A. A. (2009). Directionality of change in youth anxiety treatment involving parents: An initial examination. Journal of Consulting and Clinical Psychology, 77, 474–485. Skinner, E. A. (1995) Percieved control, motivation, and coping. Thousand Oaks, CA: Sage. Skinner, E. A., & Edge, K. (1998). Reflections on coping and development across the lifespan. International Journal of Behavioral Development, 22, 357–366. doi: 10.1080/016502598384414 Skinner, E. A., & Edge, K. (2002a). Parenting, motivation, and the development of children's coping. In R. A. Dienstbier & L. J. Crockett (Eds.), Agency, motivation, and the life course: Vol. 48 of the Nebraska Symposium on Motivation (pp. 77–143). Lincoln, NE: Nebraska University Press.

Skinner, E. A., & Edge, K. (2002b). Self-determination, coping, and development. In E. L. Deci & R. M. Ryan (Eds.), Handbook of self-determination research (pp. 297–337). Rochester, NY: University of Rochester Press. Skinner, E. A., Edge, K., Altman, J., & Sherwood, H. (2003). Searching for the structure of coping: A review and critique of category systems for classifying ways of coping. Psychological Bulletin, 129, 216–269. doi: 10.1037/0033-2909.129.2.216 Skinner, E. A., Johnson, S. J., & Snyder, T. (2005). Six dimensions of parenting: A motivational model. Parenting: Science and Practice, 2, 175–235. Skinner, E. A., Pitzer, J. R., & Steele, J. (2013). Coping as part of motivational resilience in school: A multi-dimensional measure of families, allocations, and profiles of academic coping. Journal of Educational and Psychological Measurement, 73, 803–835. Skinner, E. A., & Wellborn, J. G. (1994). Coping during childhood and adolescence: A motivational perspective. In D. L. Featherman, R. M. Lerner, & M. Perlmutter (Eds.), Lifespan development and behavior, Vol. 12. (pp. 91–133). Hillsdale, NJ: Erlbaum. Skinner, E. A., & Zimmer-Gembeck, M. J. (2007). The development of coping. Annual Review of Psychology, 58, 119–144. doi: 10.1146/annurev.psych.58.110405.085705 Skinner, E. A., & Zimmer-Gembeck, M. J. (2009). Challenges to the developmental study of coping. New Directions in Child and Adolescent Development, 2009, 5–17. doi: 10.1002/cd.239 Skinner, E. A., & Zimmer-Gembeck, M. J. (2011). Perceived control and the development of coping. In S. Folkman (Ed.), Oxford handbook of stress, health, and coping (pp. 35–59). New York, NY: Oxford University Press. Skinner, E. A., & Zimmer-Gembeck, M. J. (in press). The development of coping: Stress, neurophysiology, social relationships and resilience during childhood and adolescence. New York, NY: Springer. Smith, C. L., Calkins, S. D., & Keane, S. P. (2006). The relation of maternal behavior and attachment security to toddlers' emotions and emotion regulation. Research in Human Development, 3, 21–-31. Smith, C. L., Eisenberg, N., Spinrad, T. L., Chassin, L., Morris, A. S., Kupfer, A.,… & Kwok, O. (2006). Children's coping strategies and coping efficacy: Relations to parent socialization, child adjustment, and familial alcoholism. Development and Psychopathology, 18, 445–469. doi: 10.1017/S095457940606024X Soper, A. C., Wolchik, S. A., Tein, J.-Y., & Sandler, I. N. (2010). Mediation of a preventive intervention's 6-year effects of health risk behaviors. Psychology of Addictive Behaviors, 24, 300–310.

Sorce, J. F., Emde, R. N., Campos, J., & Klinnert, M. D. (1985). Maternal emotional signaling: Its effect on the visual cliff behavior of 1-year-olds. Developmental Psychology, 21, 195–200. Sorgen, K. E., & Manne, S. L. (2002). Coping in children with cancer: Examining the goodness-of-fit hypothesis. Children's Health Care, 31, 191–207. doi: 10.1207/S15326888CHC3103_2 Spangler, G., & Grossman, K. E. (1993). Biobehavioral organization in securely and insecurely attached infants. Child Development, 64, 1439–1450. Spangler, G., & Schieche, M. (1998). Emotional and adrenocortical responses of infants to the strange situation: The differential function of emotional expression. International Journal of Behavioral Development, 22, 681–706. doi: 10.1080/016502598384126 Spangler, G., Schieche, M., Ilg, U., Maier, U., Ackerman, C. (1994). Maternal sensitivity as an external organizer for biobehavioral regulation in infancy. Developmental Psychobiology, 27, 425–437. Spear, L. P. (2003). Adolescent brain development and animal models. Annals of the New York Academy of Sciences, 1021, 23–26. Spencer, M. B. (2006). Penomenology and ecological systems theory: Development of diverse groups. In W. Damon & R. Lerner (Eds.), Handbook of child psychology, Vol. 1: Theoretical models of human development (6th ed., pp. 829–893). Hoboken, NJ: Wiley. Sroufe, L. A. (1996). Emotional development: The organization of emotional life in the early years. New York, NY: Cambridge University Press. Sroufe, L. A. (2007). The place of development in developmental psychopathology. In Multilevel dynamics in developmental psychopathology: The Minnesota Symposia on Child Psychology (Vol. 34, pp. 285–299). Mahwah, NJ: Lawrence Erlbaum. Sroufe, L. A. (2009). The concept of development in developmental psychopathology. Child Development Perspectives, 3, 178–183. Sroufe, L. A., Egeland, B., Carlson, E., & Collins, W. A. (2005). The development of the person: The Minnesota Study of Risk and Adaptation from Birth to Adulthood. New York, NY: Guilford Press. Sroufe, L. A., & Waters, E. (1977). Attachment as an organizational construct. Child Development, 48, 1184–1199. Stansbury, K., & Gunnar, M. R. (1994). Adrenocortical activity and emotion regulation. Monographs of the Society for Research in Child Development, 59(2–3), 108–134. doi: 10.1111/j.1540-5834.1994.tb01280.x Stanton, A. L., Danoff-Burg, S., Cameron, C. L., & Ellis, A. P. (1994). Coping through emotional approach: Problems of conceptualization and confounding. Journal of Personality

and Social Psychology, 78, 1150–1169. doi: 10.1037/0022-3514.66.2.350 Steinberg, L., Dahl, R., Keating, D., Kupfer, D. J., Masten, A. S., & Pine, D. S. (2006). The study of developmental psychopathology in adolescence: Integrating affective neuroscience with the study of context. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology: Vol. 2. Developmental neuroscience (pp. 710–741). Hoboken, NJ: Wiley. Steinberg, L., & Morris, A. S. (2001). Adolescent development. Annual Review of Psychology, 52, 83–110. Stewart, S. M., Betson, C., Lam, T. H., Marshall, I. B., Lee, P. W. H., & Wong, C. M. (1997). Predicting stress in first year medical students: a longitudinal study. Medical Education, 31, 163–168. doi: 10.1111/j.1365-2923.1997.tb02560.x Sugimura, N., Rudolph, K. D., & Agoston, A. M. (2014). Depressive symptoms following coping with peer aggression: Moderating role of negative emotionality. Journal of Abnormal Child Psychology, 42, 563–575. Suveg, C., Sood, E., Comer, J. S., & Kendall, P. C. (2009). Changes in emotion regulation following cognitive-behavioral therapy for anxious youth. Journal of Clinical Child and Adolescent Psychology, 38, 390–401. Taylor, S. E., & Stanton, A. L. (2007). Coping resources, coping processes, and mental health. Annual Review of Clinical Psychology, 3, 377–401. doi: 10.1146/annurev.clinpsy.3.022806.091520 Tedeschi, R. G., Park, C. L., & Calhoun, L. G. (Eds.). (1998). Posttraumatic growth: Positive changes in the aftermath of crisis. New York, NY: Psychology Press. Tein, J.-Y., Sandler, I. N., Ayers, T. S., & Wolchik, S. A. (2006). Mediation of the effects of the Family Bereavement Program on mental health problems of bereaved children and adolescents. Prevention Science, 7, 179–195. Tein, J.-Y., Sandler, I. N., MacKinnon, D. P., & Wolchik, S. A. (2004). How did it work? Who did it work for? Mediation in the context of a moderated prevention effect for children of divorce. Journal of Consulting and Clinical Psychology, 72, 617–624. Thompson, R. A. (1990). Emotion and self-regulation. In R. A. Thompson (Ed.), Socioemotional development. Nebraska symposium on motivation (Vol. 36, pp. 383–483). Lincoln, NB: University of Nebraska Press. Thompson, R. A., & Meyer, S. (2007). The socialization of emotion regulation in the family. In J. J. Gross (Ed.), Handbook of emotion regulation (pp. 249–268). New York, NY: Guilford Press. Thompson, S. F., Zalewski, M., & Lengua, L. J. (2014). Appraisal and coping styles account for the effects of temperament on preadolescent adjustment. Australian Journal of Psychology,

66, 122–129. doi:10.111/ajpy.12048 Timko, C., Moos, R. H., & Michelson, D. J. (1993). The contexts of adolescents' chronic life stressors. American Journal of Community Psychology, 21, 397–420. doi: 10.1007/BF00942150 Tolan, P., & Grant, K. (2009). How social and cultural contexts shape the development of coping: Youth in the inner city as an example. New Directions in Child and Adolescent Development, 2009, 124, 61–74. 10.1002/cd.242 Tolan, P. H., Gorman-Smith, D., Henry, D., Chung, K. S., & Hunt, M. (2002). The relation of patterns of coping of inner-city youth to psychopathology symptoms. Journal of Research on Adolescence, 12, 423–449. doi: 10.1111/1532-7795.00040 Treynor, W., Gonzalez, R., & Nolen-Hoeksema, S. (2003). Rumination reconsidered: A psychometric analysis. Cognitive Therapy and Research, 27(3), 247–259. Tsujimoto, S. (2008). The prefrontal cortex: Functional neural development during early childhood. Neuroscientist, 14(4), 345–358. doi: 10.1177/1073858408316002 Uhlhaas, P. J., Pipa, G., Lima, B., Melloni, L., Neuenschwander, S., Nikolić, D.,… & Singer, W. (2009). Neural synchrony in cortical networks: History, concept and current status. Frontiers in integrative neuroscience, 3, 17. doi: 10.3389/neuro.07.017.2009 Umaña-Taylor, A. J., & Updegraff, K. A. (2007). Latino adolescents' mental health: Exploring the interrelations among discrimination, ethnic identity, cultural orientation, self-esteem, and depressive symptoms. Journal of Adolescence, 30, 549–567. doi: 10.1016/j.adolescence.2006.08.002 Valiente, C., Fabes, R. A., Eisenberg, N., & Spinrad, T. L. (2004). The relations of parental expressivity and support to children's coping with daily stress. Journal of Family Psychology, 18, 97–106. Valiente, C., Lemery-Chalfant, K., & Reiser, M. (2007). Pathways to problem behaviors: Chaotic homes, parent and child effortful control, and parenting. Social Development, 16, 249–267. Van De Ven, M. O., & Engels, R. C. (2011). Quality of life of adolescents with asthma: The role of personal, coping strategies, and symptom reporting. Journal of Psychosomatic Research, 71, 166–173. Vélez, C. E., Wolchik, S. A., Tein, J., & Sandler, I. (2011). Protecting children from the consequences of divorce: A longitudinal study of the effects of parenting on children's coping processes. Child Development, 82, 244–257. Wadsworth, M. E., & Berger, L. E. (2006). Adolescents coping with poverty-related family stress: Prospective predictors of coping and psychological symptoms. Journal of Youth and

Adolescence, 35, 54–67. doi: 10.1007/s10964-005-9022-5 Wadsworth, M. E., & Compas, B. E. (2002). Coping with family conflict and economic strain: The adolescent perspective. Journal of Research on Adolescence, 12, 243–274. doi: 10.1111/1532-7795.00033 Walker, L. S., Smith, C. A., Garber, J., & Claar, R. L. (2007). Appraisal and coping with daily stressors by pediatric patients with chronic abdominal pain. Journal of Pediatric Psychology, 32, 206–216. doi: 10.1093/jpepsy/jsj124 Waters, A. M., Donaldson, J., & Zimmer-Gembeck, M. J. (2008). Cognitive behavioural therapy combined with an interpersonal skills component in the treatment of generalized anxiety disorder in adolescent females: A case series. Behavior Change, 25, 35–43 Watson, D., Kotov, R., & Gamez, W. (2006). Basic dimensions of temperament in relation to personality and psychopathology. In R. F. Krueger (Ed.), Personality and psychopathology (pp. 7–38). New York, NY: Guilford Press. Watson, J. S., & Ramey, C. T. (1972). Reactions to response-contingent stimulation in early infancy. Merrill-Palmer Quarterly, 18, 219–227. Webb, T. L., Miles, E., & Sheeran, P. (2012). Dealing with feeling: A meta-analysis of the effectiveness of strategies derived from the process model of emotion regulation. Psychological Bulletin, 138, 775–808. doi: 10.1037/a0027600 Weems, C. F., & Silverman, W. K. (2006). An integrative model of control: Implications for understanding emotion regulation and dysregulation in childhood anxiety. Journal of Affective Disorders, 91, 113–124. doi: 10.1016/j.jad.2006.01.009 Wei, M., Heppner, P. P., & Mallinckrodt, B. (2003). Perceived coping as a mediator between attachment and psychological distress: A structural equation modeling approach. Journal of Counseling Psychology, 50, 438–447. Weiner, B. (1985). An attributional theory of achievement motivation and emotion. Psychological Review, 92, 548–573. doi: 10.1037/0033-295X.92.4.548 Weisz, J. R., Hawley, K. M., & Jensen Doss, A. (2004). Empirically tested psychotherapies for youth internalizing and externalizing problems and disorders. Child and Adolescent Psychiatric Clinics of North America, 13, 729–816. Weisz, J. R., Thurber, C. A., Sweeney, L., Proffitt, V. D., & LeGagnoux, G. L. (1997). Brief treatment of mild-to-moderate child depression using primary and secondary control enhancement training. Journal of Consulting and Clinical Psychology, 65, 703–707. Werner, E. E. (1993). Risk, resilience, and recovery: Perspectives from the Kauai longitudinal study. Development and Psychopathology, 5, 503–515. doi: 10.1017/S095457940000612X White, R. W. (1974). Strategies for adaptation: An attempt at systematic description. In G. V

Coelho, D. A. Hamburg, & J. E. Adams (Eds.), Coping and adaptation (pp. 47–68). New York, NY: Basic Books. Windle, M. (1992). A longitudinal study of stress buffering for adolescent problem behaviors. Developmental Psychology, 28, 522–530. doi: 10.1037/0012-1649.28.3.522 Wolchik, S., & Sandler I. (Eds.). (1997). Handbook of children's coping: Linking theory and intervention. New York, NY: Plenum Press. Wolchik, S. A., West, S. G., Sandler, I. N., Tein, J.-Y., Coatsworth, D., Lengua, L.,… &Griffin, W. A. (2000). An experimental evaluation of theory-based mother and mother-child programs for children of divorce. Journal of Consulting and Clinical Psychology, 68, 843–856. doi: 10.1037/0022-006X.68.5.843 Zelazo, P. D., Muller, U., Frye, D., & Marcovitch, S. (2003). The development of executive function in early childhood. Monographs of the Society for Research on Child Development, 68, vii–137. Zelazo, P. D., Reznick, J. S., & Pinon, D. (1995) Response control and the execution of verbal rules. Developmental Psychology, 31, 508–517. doi: 10.1037/0012-1649.31.3.508 Zeman, J., Cassano, M., Perry-Parrish, C., & Stegall, S. (2006). Emotion regulation in children and adolescents. Journal of Developmental & Behavioral Pediatrics, 27, 155–168. Zimmer-Gembeck, M. J., & Helfand, M. (2008). Ten years of longitudinal research on U.S. adolescent sexual behavior: The evidence for multiple pathways to sexual intercourse, and the importance of age, gender and ethnic background. Developmental Review, 28, 153–224. Zimmer-Gembeck, M. J., Hunter, T. A., & Pronk, R. (2007). A model of behaviors, peer relations and depression: Perceived social acceptance as a mediator and the divergence of perceptions. Journal of Social and Clinical Psychology, 26, 273–302. doi: 10.1521/jscp.2007.26.3.273 Zimmer-Gembeck, M. J., Lees, D. C., Bradley, G. L., & Skinner, E. A. (2009). Use of an analogue method to examine children's appraisals of threat and emotion in response to stressful events. Motivation and Emotion, 33, 136–149. doi: 10.1007/s11031-009-9123-7 Zimmer-Gembeck, M. J., Lees, D., & Skinner, E. A. (2011). Children's emotions and coping with interpersonal stress as correlates of social competence. Australian Journal of Psychology, 63, 131–141. doi: 10.1111/j.1742-9536.2011.00019.x Zimmer-Gembeck, M. J., & Nesdale, D. (2013). Anxious and angry rejection sensitivity, social withdrawal, and retribution in high and low ambiguous situations. Journal of Personality, 81, 29–38. doi: 10.1111/j.1467-6494.2012.00792.x Zimmer-Gembeck, M. J., Nesdale, D., McGregor, L., Mastro, S., Goodwin, B., & Downey, G. (2013). Comparing reports of peer rejection: Associations with rejection sensitivity,

victimization, aggression, and friendship, Journal of Adolescence, 36, 1237–1246. Zimmer-Gembeck, M. J., Siebenbruner, J., & Collins, W. A. (2001). Diverse aspects of dating: Associations with psychosocial functioning from early to middle adolescence. Journal of Adolescence, 24, 313–336. doi: 10.1006/jado.2001.0410 Zimmer-Gembeck, M. J., Siebenbruner, J., & Collins, W. A. (2004). A prospective study of intraindividual and peer influences on adolescents' heterosexual romantic and sexual behavior. Archives of Sexual Behavior, 33, 381–394. Zimmer-Gembeck, M. J., & Skinner, E. A. (2011). The development of coping across childhood and adolescence: An integrative review and critique of research. International Journal of Behavioral Development, 35, 1–17. doi: 10.1177/0165025410384923 Zimmer-Gembeck, M. J., Skinner, E. A., Morris, H., & Thomas, R. (2013). Anticipated coping with interpersonal stressors: Links with the emotional reactions of sadness, anger, and fear. Journal of Early Adolescence, 33, 684–709. doi: 10.1177/0272431612466175

Chapter 11 Temperament and Developmental Psychopathology Cynthia Stifter and Jessica Dollar This chapter was written with support from grants funded by the NIDDK (DK081512) and the John Templeton Foundation (#43177) awarded to the first author. INTRODUCTION WHAT IS TEMPERAMENT? The Principles of Temperament THE STRUCTURE OF TEMPERAMENT Higher-Order Temperament Factors Higher-Order versus Lower-Order Traits A Categorical Approach to Temperament TEMPERAMENT AND PERSONALITY THE MEASUREMENT OF TEMPERAMENT Questionnaire-Based Assessments of Temperament Observational Measures of Temperament Other Objective Measures of Temperament Multimethod Approach to Measuring Temperament TEMPERAMENT AND DEVELOPMENTAL PSYCHOPATHOLOGY Internalizing and Externalizing Problem Behaviors TEMPERAMENT AND PSYCHOPATHOLOGY: REVIEW OF THE RESEARCH LITERATURE Negative Emotionality Surgency Positive Emotionality Activity Level Effortful Control TEMPERAMENT × TEMPERAMENT INTERACTIONS Effortful Control × Temperament

Surgency × Temperament Positive Emotionality × Temperament NEUROBIOLOGY LINKING TEMPERAMENT AND DEVELOPMENTAL PSYCHOPATHOLOGY Cardiac Measures Electrical Brain Activity Functional Neuroimaging Psychobiological Measures of Stress Genetic Measures FINAL REMARKS: SUMMARY AND FUTURE DIRECTIONS TRANSLATIONAL IMPLICATIONS CONCLUSION REFERENCES

Introduction The field of developmental psychopathology has relied extensively on the study of normal development in order to understand psychological maladjustment (Cicchetti & Cohen, 2006; Sroufe, 2009). Along with current models of development that consider multiple determinants of developmental outcomes, developmental psychopathology is concerned with understanding how both child and environmental factors predict the development of problem behavior in children. And, consistent with research in adult psychopathology, research in developmental psychopathology has examined how individual differences in emotionality, attention/regulation and activity, better known as temperament, have contributed both directly and indirectly to behavioral disorders. Although there is a long tradition behind the concept of temperament and its role in psychopathology, it was not until the twentieth century that individual variation in child behavior became the focus of study. And it was not until the early 1960s that such variation became the focus of child psychologists and psychiatrists interested in understanding the origins of children's behavioral problems (Thomas & Chess, 1977). Since that time, various developmental models of temperament have been proposed leading to a vast array of instruments and observational methods for assessing temperament in children, all in the service of understanding the defining characteristics and structure of temperament as well as its impact on other developmental outcomes, including psychopathology. In this chapter we review the current literature linking infant and child temperament to developmental psychopathology. Our review is an update to the reviews found in previous editions of this volume provided by Rothbart and colleagues in 1995 and in 2006 (Rothbart &

Posner, 2006; Rothbart, Posner, & Hershey, 1995). We concentrate on studies published after 2000, a period of increased study in this area (Zentner & Shiner, 2012). Paralleling this wave of new research was a growing understanding of the biological systems underlying both temperament and psychopathology (Fox, Henderson, Pérez-Edgar, & White, 2008). Likewise, temperament theory and research was going through a renaissance of sorts with an update of the definition of temperament and introduction of new approaches to the study of temperament. Hence, another goal of this chapter is to review these exciting changes. Before the review of the most current research in temperament and developmental psychopathology, we introduce the concept of temperament and consider a number of agreed-on principles. We follow this with a discussion of the structure of temperament, including a review of the contemporary theoretical models, as well an examination of the more traditional dimensional approach and the more recent (resurgent) categorical approach to the structure of temperament. We then consider the measurement of temperament by questionnaire, laboratory/home observation, and other methods. We conclude our introduction by discussing the proposed links between temperament and developmental psychopathology. Following our review of the current research on temperament and developmental psychopathology, we provide a brief overview of the biological underpinnings of temperament and their links to developmental psychopathology. Based on our review of the research literature, we conclude the chapter with directions for future research.

What Is Temperament? In 1987, several prominent and active researchers in the field of temperament gathered for a roundtable sponsored by the Society for Research in Child Development (Goldsmith et al., 1987) to discuss the current state of the field. Although there was significant consensus, the participants' conceptualizations differed in several ways depending on what they chose to emphasize in their models. For example, A. Buss and Plomin (1984) emphasized the genetic basis of temperament while Goldsmith and Campos's framework (1982) was emotion based. In his commentary on the roundtable, McCall attempted a definition that integrated the agreed-on concepts: “temperament consists of relatively consistent, basic, dispositions that underlie and modulate the expression of activity, reactivity, emotionality and sociability” (Goldsmith et al., 1987, p. 524). More recently, researchers who were affiliated with each of the original roundtable participants met in the same format 25 years later to review what was learned about temperament in the interim (Shiner et al., 2012). Based on their own research and knowledge of the current literature on child temperament, the progeny of the original roundtable participants updated the definition of temperament as “early emerging basic dispositions in the domains of activity, affectivity, attention, and self-regulation, and these dispositions are the product of complex interactions among genetic, biological, and environmental factors across time” (Shiner et al., 2012, p. 437). As can be seen, this definition was expanded to illustrate that temperament is not a static construct but one that develops. This definition of temperament is consistent with the six key criteria for temperament identified by Zentner and Bates (2008). (Zentner & Bates, 2008). Several of the criteria overlap with the

preceding definitions of temperament (domains of affect, activity and attention; early appearance; linked to biological mechanisms; relatively enduring). Additional criteria included: (1) variations in expression as reflected in response intensities, latencies, durations, thresholds, and recovery times; (2) comparability with individual differences observed in other social mammals; and (3) predictions to “conceptually coherent outcomes,” such as internalizing and externalizing problem behaviors. These current definitions/criteria of temperament are important as they provide a consensual framework to guide future research and allow for the comparison across studies on child temperament.

The Principles of Temperament The contemporary definitions of temperament strive to include the many principles of temperament on which the different models agree (Goldsmith et al., 1987). It is worth discussing a few of these principles in light of the following review of the literature linking temperament to developmental psychopathology. The first principle is that temperament reflects individual differences that are relatively stable. Stability, which comes in many forms (e.g., homotypic, heterotypic; Caspi, 1998), applies theoretically to temperament, given the characteristics used to define temperament such as “basic,” “disposition,” and “enduring.” That temperament is stable also increases its predictability. That is, parents may expect their child to respond similarly across situations. Likewise, when asked to describe their children, parents use descriptions that reflect stable characteristics, such as “She is outgoing” and “He is shy,” which are easily understood by others (Kohnstamm, Mervielde, Besevegis, & Halverson, 1995). Despite this notion that temperament is stable, the qualifier “relatively” has been added in current definitions of infant and child temperament, reflecting an understanding that temperament traits may change over time, given biological maturation and environmental input. Evidential support for relative stability of temperament is found in studies that show the consistency of temperament to be low to moderate with stability coefficients ranging from zero (0) to around .50, depending on how temperament is measured as well as the ages over which stability was calculated (Durbin, Hayden, Klein, & Olino, 2007; Lemery, Goldsmith, Klinnert, & Mrazek, 1999; Roberts & DelVecchio, 2000). Greater consistency is usually found for parent ratings, older children (>3 years of age), and across shorter lengths of time while lower stability coefficients are typically found for laboratory/observational measures, infants and toddlers, and across longer periods of time. Further evidence is provided by quantitative behavior genetic studies. High heritability of a temperament trait would indicate excellent stability, and for many traits, heritability effects are robust. However, genetic effects across studies range from .20 to .60, and there is little consistency across studies examining the same dimension. Taken together, these data suggest the possibility of environmental effects as well as error in measurement (Saudino & Wang, 2012). Recent research examining the genetic influences on continuity and change in temperament from infancy to preschool revealed that the modest continuity of several observed traits was due to genetic factors, but other dimensions, such as shyness, were influenced by genetic and environmental factors (Saudino & Cherny, 2001). Although these studies indicate the genetic bases for stability of temperament, it is important to note that the average age-to-age correlation was .24, confirming that stability of

temperament is modest, particularly in early childhood. A second principle of temperament is that temperament develops. This is an often-neglected principle given the folk theories about the endogenous nature and stability of temperament. However, the modest correlations of temperament across age suggest that change in temperament is more likely to occur; this may be due to maturation of biological systems, the emergence of other temperament dimensions, as well as input from the environment, particularly parents. Many of the dimensions found within contemporary temperament models can be said to develop. Consider Rothbart and Derryberry's (1981) definition of temperament; they define temperament as individual differences in reactivity and regulation. Although several forms of reactivity are observable at birth, its expression is undifferentiated in terms of emotion (e.g., distress) and diffuse with relation to motor activity. With development, different emotions emerge and consequently reactivity becomes more specific. For example, anger reactivity and social smiling emerge around 3 months of age while fear emerges after 6 months of age and is dependent on the start of self-locomotion (Barrett & Campos, 1987). Studies examining the emergence of fear-based inhibition support this later emergence of temperamental fear, demonstrating that prior to 6 months, infants approach low-intensity (e.g., plate, saucer) and high-intensity toys (e.g., flashing light) with similar latencies but by 12 months of age, latencies to approach lengthen to high-intensity toys, and individual differences in latencies to approach are predictive of later inhibited/uninhibited temperament (Putnam & Stifter, 2002; Rothbart, 1988). Regulation, in contrast, begins to develop at the end of the first year, although attentional behaviors, such as focusing and sustained attention that form the foundation of self-regulation (Garon, Bryson, & Smith, 2008), can be observed earlier. By the end of the first year, infants are demonstrating rudimentary self-regulation or effortful control, as demonstrated by their ability to initiate, maintain, and cease activities and their awareness of rules (Kopp, 1982). Throughout the next 2 years, the ability to control emotions, behavior, and cognition develops rapidly, moving from voluntary inhibition of behavior to the ability to suppress a dominant response in favor of a nondominant response (Kochanska, Coy, & Murray, 2001). Likewise, executive control of attention, which is linked to activation of the anterior cingulate cortex (ACC) (Rudea, 2012) emerges during this period. The emergence of temperamentally based self-regulation or effortful control overlaps significantly with the development of executive functions, a rubric that subsumes cognitive processes that control thought and action including planning, working memory, set shifting, error detection, and inhibitory control of prepotent responses (Blair, Zelazo, & Greenberg, 2005; Welsh, Pennington, & Groisser, 1991). Thus, both temperamental reactivity and regulation develop across infancy and early childhood. Indeed, while specific forms of reactivity emerge within the first year, development continues across childhood such that levels of reactivity typically decrease due to increases in self-regulation. Whether these changes are due to the structure of temperament or its expression continues to be debated. What is known is that the inability to regulate emotional reactivity is related to risk for behavior problems and is symptomatic of child clinical disorders. It is important to note that self-regulation refers to voluntary, executive control, which is

contrasted with what Rothbart (Rothbart & Bates, 2006) calls “passive control” and Eisenberg (Eisenberg & Morris, 2002; Eisenberg et al., 2004) calls “reactive control.” Although these terms appear to be opposing, they are quite similar in that both terms define nonvoluntary or automatic control that is driven by approach/withdrawal processes. Rothbart's term is focused exclusively on fear-based behavioral control, such as when inhibited children withdraw from an object/person they perceive as a potential threat. Thus, their fear constrains or regulates their behavior. Interestingly, passive control is related to later effortful control such that fearful children exhibit better effortful control than fearless children (Kochanska & Knaack, 2003). Eisenberg extended Rothbart's term to include impulsive approach as a form of reactive control in that impulsive children are fueled by their approach system without any voluntary regulation. Importantly, both forms of reactive control have been related to problem behaviors (see below). Acknowledging that temperament develops may be particularly important for understanding how temperament influences developmental psychopathology. The central tenet of developmental psychopathology is that behavioral and emotional disorders are the result of a complex developmental process (Sroufe & Rutter, 1984). Research on this process requires prospective longitudinal study that starts in infancy and includes genetic, physiological, contextual, and experiential components and their development over time (Sroufe, 2009). Assessing the development of the emotional and regulatory components of temperament is essential to identifying the pathways to and away from psychopathology. For example, understanding when children develop regulatory strategies in light of their temperamental reactivity and how effective these strategies are developmentally will assist in the assessment of anger reactivity as a symptom of psychological maladjustment rather than the extreme end of a temperamental characteristic. A third principle to which all contemporary temperament models subscribe is that temperament interacts with the environment to affect the child's development. This principle is consonant with the premise that interactions between children and their parents are bidirectional (Bell, 1968) and transactional (Sameroff, 2009) such that the characteristics of children, including temperament, contribute to their own development through interactions with the parenting environment across time. Children's temperament can, for example, evoke certain responses from parents, which can, in turn, affect how children consequently respond. Studies have shown how specific parenting behaviors interact with temperament either to attenuate or exacerbate later reactivity as well as to support or undermine the development of selfregulation, which has been shown to have implications for the development of psychopathology (Kiff, Lengua, & Zalewski, 2011). Several theoretical approaches attempt to explain how temperament by environment interactions impact developmental outcomes, including psychopathology. The organismic specificity hypothesis (Gandour, 1989; Wachs, 1991) proposes that similar environments will affect children differently due to inherent individual differences in each child. Thus, a highly negative infant will respond to parental soothing differently from an infant who is low in negativity. This differential response may impact the development of the child, depending on how the parent adjusts his or her behavior accordingly. This model of development parallels

the diathesis-stress or dual-risk model utilized to describe the development of psychopathology (Ingram & Luxton, 2005). More specifically, the diathesis-stress model views individuals who possess certain vulnerabilities that serve as a diathesis to increase their risk for negative outcomes (e.g., psychopathology) when combined with adverse environments while those who are not characterized by vulnerabilities are unaffected by the same environments and thus considered resilient (Luthar, 2006). In other words, some children are more strongly affected by poor environmental conditions, whereas other children appear to be more resilient to difficult environmental situations; this resilience frequently is believed to be a result of the child's temperament. A child who is temperamentally fearful, for example, and who has multiple nonparental caregivers with little predictability may develop general anxiety disorder while a less fearful child in the same circumstance would be a low risk for such a disorder (Degnan, Almas, & Fox, 2010). An important aspect of this model is that all children, including both those who are resilient and those who are vulnerable, thrive similarly in supportive or nonmaladaptive environments. Some researchers have recently argued that most research in developmental psychopathology has not considered the full range of environments or outcomes, ranging from positive to negative (van IJzendoorn & Bakermans-Kranenburg, 2012a). Two very similar developmental models have been put forward to address this gap: the differential susceptibility hypothesis (Belsky & Pluess, 2009) and the biological sensitivity to context theory (Boyce & Ellis, 2005). These models propose that certain individuals are susceptible to both “risk-promoting and development-enhancing environmental conditions” and that this susceptibility is found within the individual's biology (B. J. Ellis, Boyce, Belsky, Bakermans-Kranenburg, & van IJzendoorn, 2011). In other words, susceptible individuals are not only more affected by adverse environments than less susceptible individuals but are also more open to positive, enriching environments that will affect outcomes for these individuals above and beyond those individuals who are less susceptible (Belsky, 1997; Belsky & Pluess, 2009). Temperament, particularly difficult temperament or high levels of negative emotionality, and the biological systems and genetics that underlie temperament, have been used as examples of a susceptibility factor. Van Zeijl and colleagues (2007), for example, found that children with difficult temperaments were more susceptible to both negative and positive discipline than were children with easy temperaments in predicting the highest and lowest levels of externalizing problems, respectively. In another study, Kim and Kochanska (2012) demonstrated that infants with high negative emotionality exhibited the lowest self-regulation when their mothers were not mutually responsive and the highest levels of self-regulation when their mothers were very responsive. Biological sensitivity might be exemplified by activity in the parasympathetic nervous system (PNS) as measured by respiratory sinus arrhythmia (RSA). Findings from a study on marital conflict showed that when children with low RSA live in a household high in marital conflict, they are most anxious; when children with low RSA live in households with low marital conflict, they are least anxious (El-Sheikh, Hager, & Whitson, 2001). It is important to note that fearful children are characterized by low RSA, suggesting that their propensity for anxiety disorder would not only be lowered under positive conditions but buffered to the extent that they are the least anxious among other fearful and nonfearful

children. As many of the proposed pathways to developmental psychopathology include the role of the environment, these developmental models become increasingly important in understanding the relationship between temperament and psychopathology as well as in guiding future research. Moreover, the differential susceptibility and biological sensitivity models of person by environment interactions have only recently been fully tested because they are relatively new. However, several studies show the pattern suggested by these models. In the Temperament and Psychopathology: A Review of the Literature section we note which studies are supportive of temperament as a differential susceptibility in the development of child psychopathology.

The Structure of Temperament Conceptually and empirically, the structure and measurement of temperament in childhood has traditionally been dimensional. This is likely because the predominant conceptualization of adult personality/temperament that influenced the consideration of temperament earlier in life was dimensional, and that measurement of personality in adults and temperament in children was and is based primarily on questionnaires. In this chapter, we focus on the four contemporary models of temperament that dominate the field: Thomas and Chess's (1968) behavioral-style model, A. Buss and Plomin's (1975) genetic criteria model, Rothbart and colleagues' (1981) psychobiological model, and Goldsmith and Campos's (1982) emotion-based model. Drawn from the dimensions that emerged from the interviews of Thomas and Chess and subsequent theory-driven and factor analytic work on child temperament, researchers converged around defining temperament as individual differences in three major areas: emotions, activity, and attention/regulatory behaviors (see Mervielde & De Pauw, 2012; Zentner & Bates, 2008, for a more expanded list of converging traits). Table 11.1 lists the specific dimensions of the four child temperament approaches that correspond with these three broad temperament factors as well as those dimensions that are specific to the temperament model. As noted, each theory emphasizes different aspects or criteria for temperament as well as the different developmental ages that each model covers. Both of these factors influenced the type and number of dimensions proposed by each theory. Here we describe briefly the dimensions from each temperament model that fall within these broad factors. In addition, these broad temperament factors are used later in the chapter to guide the organization of our review of the current research on the links between temperament and developmental psychopathology.

Table 11.1 Dimensions of contemporary developmental models of temperament Thomas & Chess A. Buss & (1968) Plomin (1975)

Rothbart et al. (1981) Goldsmith Infant Child & Campos (1982)

Emotionality Intensity of Emotionality Anger reaction; Fear Threshold of Smiling response; Quality of mood Activity Level Activity level Activity Activity level level Attention Behaviors Attention span Interest Regulation Approach/ Sociabilty withdrawal

Distractibility Other dimensions Rhythmicity; Adaptability

Impulsivity

Distress to limitations; Fear; Sadness; Low/high intensity pleasure; Smiling and laughter

Anger/frustration; Fear; Sadness; Low/high intensity pleasure; Smiling and laughter; Discomfort

Activity level

Activity level

Duration of orienting

Attentional focusing

Approach

Approach; Shyness

Falling reactivity; Soothability

Falling reactivity and soothability Inhibitory control Impulsivity

Cuddliness; Perceptual Perceptual sensitivity sensitivity; Vocal reactivity

The broad factor of emotionality, which describes intensive and temporal variations in emotion expression, is defined either generally or specifically, depending on the theoretical framework. For example, Thomas and Chess (1968) include children's quality of mood, which ranges from positive to negative, while A. Buss and Plomin's (1975) emotionality dimension focuses broadly on negative emotions and does not include consideration of positive emotions. Other researchers view individual differences in discrete emotions, such as happiness, anger, and fear. Rothbart's (1981) positive emotion dimension includes smiling/laughter and high/low-intensity pleasure. Smiling, as a reflection of joy, is also included in Goldsmith and Campos's (1982) framework. Similarly, individual differences in anger and fear are described

in both theories. However, these are termed distress to limitations (anger) and distress to novelty (fear) in the infant version of Rothbart's model (1981). The attention to discrete emotions in these models emphasizes the understanding that the functions of these emotions are different (Barrett & Campos, 1987) and thus individual differences in their expression may lead to different outcomes (Putnam & Stifter, 2005). Indeed, among negative emotions, some may be positively related, as was found between anger and sadness (Dyson, Olino, Durbin, Goldsmith, & Klein, 2012), while others may be unrelated, as in the case of anger and fear (Goldsmith, 1996). Interestingly, factor analytic data (Rothbart, Ahadi, Hershey, & Fisher, 2001) on parent ratings of child temperament and observational studies (Gagne, Van Hulle, Aksan, Essex, & Goldsmith, 2011) indicated that in some cases, happiness and anger are positively linked. This is consistent with Carver's (2004; Carver & Sheier, 1998) theoretical model of affect, which proposes that positive and negative emotions are generated in response to the effectiveness of approaching incentives (happiness and anger) or withdrawing from threat (relief, fear). Activity level, which reflects an individual's level and intensity of motor activity, is included in all modern theories of child temperament (A. Buss & Plomin, 1975, 1984; Rothbart & Derryberry, 1981; Thomas & Chess, 1968, 1977). Activity level is commonly viewed as such a critical aspect of temperament that it is even considered in Goldsmith and Campos's (1982) emotion-focused temperament framework. They proposed that motor activity is tied to each emotion and thus represents the child's emotional arousal (Goldsmith & Campos, 1982; Goldsmith & Campos, 1990). Its inclusion in A. Buss and Plomin's (1984) genetic criteria model is due to its moderate to high heritability in children, which has been found in both parent ratings of activity (Saudino & Eaton, 1991) as well as objective measurements, such as actigraphs (A. Wood, Saudino, Rogers, Asherson, & Kuntsi, 2007). Activity level is often associated with other temperament dimensions. For example, it is positively associated with anger and negatively associated with fear (Rothbart, Ahadi, & Evans, 2000). This finding confirms Goldsmith and Campos's (1982) proposal that activity level, at least in very young children, reflects emotional arousal. High levels of activity are also linked to high levels of approach. As we discuss later, these traits along with high-intensity pleasure and impulsivity, to name a few, make up the superfactor “surgency” identified by Rothbart (Rothbart, Ahadi, & Hershey, 1994). The third broad factor of temperament that emerges from these four theories is attention/regulatory behaviors. Regulation is the process that modulates (e.g., increases, decreases, or maintains) emotional and behavioral reactivity, and attention is foundational to this ability (Posner & Rothbart, 2000). Thomas and Chess's (1977) attention span dimension and Rothbart's (Rothbart & Derryberry, 1981) duration of orientation and attention focusing dimensions reflect individual differences in attention. Goldsmith and Campos's (1982) model includes the construct of interest, which is conceptualized as an emotion resulting from attention to stimuli and thus is included under this factor. Approach/withdrawal behaviors found in the Thomas and Chess and Rothbart models are listed under regulation because they increase positive affect by approaching or moving toward novel stimuli while withdrawal from novel stimuli decreases negative affect. Of course, who is more likely to use approach or

withdrawal behaviors will depend on their reactivity to novel stimuli. A. Buss and Plomin's (1984) sociability and Rothbart's shyness dimensions represent approach/withdrawal in social situations. As regulation is central to Rothbart's definition of temperament, this model has several dimensions not seen in the other models. Falling reactivity and soothability reflect the child's ability to recover from the height of reactivity and to be regulated by others. Inhibitory control, which is defined as the child's ability to plan activities and suppress reactivity when instructed, develops in early childhood and thus is not seen in measures of infant temperament. Finally, as can be seen in Table 11.1, distractibility and impulsivity were included under this broad factor. Although these dimensions represent the speed with which a child responds to stimuli, we conceptualize them here as reflecting the lack of regulation. Impulsivity was initially included in Buss and Plomin's framework and then removed upon findings of lower than criteria heritability. Recent findings, however, suggest that there is a significant genetic influence on impulsivity (Gagne & Saudino, 2010).

Higher-Order Temperament Factors As with the development of the Three-Factor (Eysenck, 1970) and Five-Factor (Costa & McCrae, 1976) models of personality and their numerous variations, temperament researchers have also examined how dimensions of temperament might relate in a coherent fashion. Examining the presence of higher-order factors is theoretically consistent with the proposal that temperament, like adult personality, is organized hierarchically (Zentner& Shiner, 2012). The use of higher-order temperament factors also has both methodological and interpretative advantages. As it is expected that many of the dimensions covary (e.g., activity level and high intensity pleasure) creating broad traits increases the power to predict relationships with other nontemperament factors (Putnam, Ellis, & Rothbart, 2001). Higher-order temperament factors also align quite well with the Three- and Five-Factor models of personality such that investigations of the links between temperament and personality are better executed and interpreted (Rothbart, Ahadi, et al., 2000). Likewise, since there is no unifying model of temperament, higher-order factors are useful for comparing across studies of temperament using different theoretical and methodological approaches. Rothbart and colleagues (Putnam et al., 2001; Rothbart, Ahadi, et al., 2001) have done the most extensive quantitative analysis of temperament dimensions across childhood. Using factor analysis on the 15 dimensions (see Table 11.1) measured with the Child Behavior Questionnaire (Rothbart et al., 2001), three superfactors were identified: Negativity, Surgency, and Effortful Control. Negativity reflects a general tendency toward negative affect and is composed of anger/frustration, fear, sadness and low soothability. The surgency factor, which defines an approach orientation that generates positive affect and activity, is comprised of positive loadings of high-intensity pleasure, activity level, impulsivity and approach, and a negative loading of shyness. Effortful control, which results from positive loadings of inhibitory control, attention focusing, low intensity pleasure, smiling/laughter, and perceptual sensitivity, represents the child's regulatory capacity. These factor structures were similar across ages 3 years to 7 years and replicated in a Chinese sample (Ahadi, Rothbart, & Ye, 1993; Rothbart, Ahadi, et al., 2001) and a Dutch sample (Mervielde & De Pauw, 2012). The

three superfactors subsequently emerged from questionnaires aimed at infants (Gartstein & Rothbart, 2003) and toddlers (Putnam, Gartstein, & Rothbart, 2006), demonstrating considerable coherence of these broad factors across childhood. Interestingly, these broad temperament factors map well onto adult personality factors, supporting the proposed link between temperament and personality traits (Shiner & Caspi, 2012). The three factors identified by Rothbart were all derived from parent ratings of their children's temperament. Recently researchers have attempted to identify higher-order factors using observational methodology. Using procedures from the Laboratory Assessment Battery (LabTAB; see the section on Observational Measures of Temperament for further details, Goldsmith, Reilly, Lemery, Longley, & Prescott, 1993), which was designed to measure a number of lower-order traits, such as anger, sadness, fear, activity, and inhibitory control, Dyson and colleagues (2012) took a two-stage factor analytic approach. Exploratory factor analysis identified five factors, which were confirmed with a second sample: Sociability, Positive affect/interest, Dysphoria, Fear/inhibition, and Constraint versus impulsivity. In another study, examiner ratings of child behavior across a number of cognitive and motor tasks were used to assess such temperamental characteristics as fearfulness, emotional lability, and impulsiveness (Caspi, Henry, McGee, Moffit, & Silva, 1995). Three factors emerged from the factor analysis: Lack of control, Approach, and Sluggishness, which described children who were passive, quiet, and withdrawing from novelty. When examining the relationship among these temperament factors and child behavior problems, lack of control was the most robust predictor. Similar results have been found with parent-reported effortful control. Taken together, the factors that emerged from studies using observations overlap (e.g., positive affect/interest, approach and surgency; constraint, lack of control, and effortful control) but are not identical to those that emerged from analysis of parent questionnaires.

Higher-Order versus Lower-Order Traits Although there are several advantages to using higher-order factors in temperament research, this does not diminish the benefits of a fine-grained approach examining lower-order traits. Consideration of specific temperament traits may reveal more definite relations with outcomes such as behavioral and emotional disorders in childhood. So while there is evidence that the higher-order factor of Negativity, which includes variations in fear, anger, sadness, and general distress, is related to child behavior problems, an examination of the lower-order traits of anger and fear reactivity show that they are differentially related to the broad-band dimensions of problem behavior. For example, anger is positively associated with externalizing while fear reactivity is associated with internalizing (see the Review of Literature section). This finding should not be surprising, as these emotions have very different adaptive goals and action patterns (Barrett & Campos, 1987). It is interesting to note that in the factor analysis that derived the three higher-order factors of Surgency, Negativity, and Effortful control, anger reactivity loaded equally across the Surgency and Negativity factors (Rothbart, Ahadi, et al., 2001). Surgency includes high approach and activity level, characteristics that, when restrained, may lead to frustration and anger. Indeed, anger is often referred to as an approach emotion (Carver & Harmon-Jones, 2009; He et al., 2010). The use of higher-order

temperament factors, therefore, may obscure important individual differences reflected in lower-order traits. Finally, it is important to note that the higher-order factors were derived from factor analysis of parent-rated temperament. As we discuss later, this method of assessing temperament is highly subjective, which may account for the associations found among the different types of infant and child negative emotions. Researchers interested in the links between temperament and developmental psychopathology, therefore, should be guided by the underlying emotions or behaviors defining the disorder. Researchers interested in oppositional defiant disorder (ODD), for example, which is characterized by ease of frustration, might consider focusing on anger reactivity, while those examining anxiety disorders would benefit from a more focused examination of fear reactivity.

A Categorical Approach to Temperament Although the dimensional approach to temperament is the predominant method by which this construct is described, the categorical approach to understanding individual differences in reactivity and regulation has recently come to the forefront. The emergence of advanced statistical techniques, such as latent profile analysis, has allowed researchers to more fully examine how patterns of temperament characteristics define homogeneous groups of individuals. The categorical approach to the study of temperament, however, is not new. In the second century, Galen described temperament categorically (e.g., sanguine, melancholic) based on the balance of the four humors of yellow bile, black bile, phlegm, and blood (Kagan, 1994). Interestingly, these types persisted across a number of cultures until the nineteenth century, when empiricism and later behaviorism emerged to explain individual differences. In the second half of the twentieth century, Thomas, Chess, and Birch (1968), the first developmentalists to consider temperament as important to understanding developmental psychopathology, created nine dimensions (see Table 11.1) based on interviews with parents which were then collapsed into three child temperament types. The easy type, which made up 40% of the sample, was more approaching of unfamiliar objects and persons, more positive in mood and adaptive. The difficult type described a child who withdrew from the unfamiliar, had an irritable, negative mood, and exhibited poor adaptation and comprised 10% of the sample. Children of this type were most likely to be rated as high in behavior problems. The third type, slow to warm up, comprised 15% of the sample and was most likely to exhibit withdrawal from novelty with some emotional distress and moderate adaptation. These categories were developed through qualitative analyses followed by factor analyses. The recent resurgence of research examining categories or types of temperament resulted in part from the person-oriented approach to individual functioning discussed by Block (1971) and further developed conceptually and methodologically by Magnusson and colleagues (Bergman & Magnusson, 1997). The theoretical basis of the person-oriented approach is holistic in that the person is viewed as integrated and hierarchically organized, and individual differences are investigated for patterns among the behaviors of interest (Magnusson, 1998) rather than a summation of the variables. Person-centered or typological methods aim to create subgroups of individuals with similar patterns of behaviors or traits (Kagan & Fox, 2006;

Zentner & Bates, 2008) based on the premise that members of a specific subgroup are more similar to each other than to members of a different subgroup (Bergman & Magnusson, 1997). This approach provides increased statistical power to identify complex interactional associations among multiple temperament traits (Laursen & Hoff, 2006; Magnusson & Bergman, 1990). It is important to note that temperament typologies were based on observations of behaviors rather than questionnaires, the predominant method of studies using the dimensional approach (cf. Dyson et al., 2012; Gagne, Van Hulle, et al., 2011). This contrast in methodology raises the issue of comparability across the dimensional and categorical approaches to temperament. However, the variable-centered and person-centered methods of studying temperament might be considered complementary to one another, not competing (Laursen & Hoff, 2006; Magnusson, 1998). Both approaches have utility in temperament research, are aimed at answering different research questions, and offer different methods of examining the data. The variable-oriented, dimensional approach might ask what the implications are of being high or low on a temperament characteristic for behavioral problems. For example, are high levels of anger reactivity related to aggressive behaviors? Research taking a person-oriented, typological approach, however, would consider the developmental outcomes of distinct subgroups of children with patterns of temperamental reactivity and regulation, such as how behaviorally inhibited children might develop anxiety disorders. There are several ways in which personality and temperament types might be identified (Robins, John, & Caspi, 1998): (1) typologies based on extreme groups; (2) typologies based on a bimodal distribution; (3) bivariate or two-dimensional typologies where two independent dimensions are crossed to form four groups of individuals; and (4) a multivariate approach to creating groups. At least two of these methods have been used to create temperament types. Using the extreme groups approach, Kagan (1995) and others (Fox, Henderson, Rubin, Calkins, & Schmidt, 2001) identified behaviorally inhibited children and uninhibited children based on a number of characteristics that reflected their approach/withdrawal tendencies that fell into the top 10% and bottom 10% of these behaviors. Research from these labs has shown that the two temperament types are distinct behaviorally, physiologically, and physically. Behaviorally inhibited children, for example, are characterized by low verbal and emotional spontaneity with unfamiliar adults, long latency to relax in new situations, a reluctance to take risks, blue eyes, and a narrow face (Kagan, 1994). As we discuss later in this chapter, they have a distinct physiological profile. In addition, several studies have shown strong associations between extreme behavioral inhibition, particularly if it is stable across childhood, and later internalizing disorders. It is important to note that the groups used by these researchers were not exhaustive, and similar to the Thomas and Chess's (1977) creation of temperament types, several children were left unclassified. A number of studies have employed a variety of multivariate approaches to derive temperament types. Using multivariate methods allows for the identification of types based on a configuration of a number of traits/dimensions. Cluster analysis is one such multivariate method and has been used to classify all children in a sample. For example, Caspi and Silva (1995) found five replicable types based on examiner ratings of several emotional,

approach/withdrawal, and regulatory characteristics of 3-year-olds: well adjusted, inhibited, undercontrolled, confident, and reserved. Using coded behaviors of children observed during a number of laboratory tasks designed to elicit approach/withdrawal and emotional reactivity, Putnam and Stifter (2005) identified three temperament types in toddlers using cluster analysis: exuberant, inhibited, and low reactive. The exuberant type was high on approach and showed high levels of positive affect whereas the inhibited children showed the lowest levels of approach and the highest levels of negative affect. The low reactive type was so labeled as they showed lower levels of both positive and negative affect as well as moderate levels of approach. One of the goals of this study was to distinguish among low approach that was due to high levels of fear to novelty and low approach due to low interest in novelty. Another multivariate method for identifying groups of children based on their temperament is to use a latent variable approach (e.g., latent class analysis, latent profile analysis) which classifies individuals based on probabilities of membership for each class. Using latent profile analysis based on variables coded from a risk-room procedure, Dollar, (2011) identified three temperament types similar to those of Putnam and Stifter (2005): exuberant, inhibited, and average. Importantly, this replication was conducted with a slightly older preschool sample using a different statistical technique. In another study, Loken (2004) replicated the infant reactivity groups first identified by Kagan (1994) using latent class analysis. Latent variable analyses have also been used to identify longitudinal profiles. For example, stable high versus stable low profiles across infancy and early childhood emerged for behavioral inhibition (Chronis-Tuscano et al., 2009) and exuberance (Degnan et al., 2010). These longitudinal profiles have been successful in predicting later behavioral disorders. Summary Contemporary temperament research has viewed the structure of temperament as dimensional. The dimensions identified by the four most popular models of temperament converge somewhat and coalesce around three general domains: emotionality, activity level, and attention/regulatory behaviors. Higher-order factors or superfactors have emerged through factor analysis of these temperament dimensions demonstrating some distinct associations among temperament traits. A resurgence of the categorical approach to temperament has produced a number of temperament types. The inhibited/shy type and the exuberant/social type have emerged rather consistently from studies using varied methodologies and samples. Both the dimensional and categorical approaches to the structure of temperament have their advantages, which should be considered when investigating the link between temperament and developmental psychopathology.

Temperament and Personality When discussing the definition or criteria for temperament, it is important to consider how temperament is related to personality. Personality in children of school and adolescent age has recently been studied, and the evidence suggests that children's personality is structured very similarly to adult personality. Although the approach to child temperament is considered

bottom up in that Thomas and Chess (1977) interviewed parents about their children's enduring characteristics, contemporary temperament theorists have said that they were influenced by adult personality research at the time (Goldsmith et al., 1987). A number of scholars focused on understanding the distinction between temperament and personality, and several models have emerged (McCrae et al., 2000; Rothbart, Ahadi, et al., 2000; Shiner & Caspi, 2012); however, very little empirical study has been conducted to test the conceptual link between temperament and personality (De Pauw, Mervielde, & Van Leeuwen, 2009). As a result, the boundaries between what is temperament and what is personality still are fuzzy. In general, the term personality describes individual differences that are the product of early emerging constitutionally based temperament traits in interaction with the social and physical environment and development in cognitions, self-concept, and more complex emotions, such as guilt and pride. Thus, personality incorporates traits that may be the continuation of temperament traits, though labeled differently, as well as attitudes, values, and coping mechanisms, all of which are proposed to be influenced by temperament (Rothbart, 2011; Shiner & Caspi, 2012). Importantly, these characteristics may be specific to a particular time or place (McAdams & Pals, 2006). Thus, the most direct link between temperament and personality is hypothesized to be through dispositions or traits: behaviors, thoughts, and emotions that are consistent across time and situation. In personality, these traits are referred most often as the Big Five and include Neuroticism, Extraversion, Agreeableness, Conscientiousness, and Openness to New Experiences (McCrae et al., 2000). These personality traits, which are composed of a number of lower-order dimensions, have been conceptually linked to both higher-order factors and lower-order temperament traits. Extraversion is similar to the higher-order factor of surgency, and both extraversion and surgency include components of approach and positive emotionality. Neuroticism, which reflects individual differences in anxiety, fearfulness, moodiness, and low frustration tolerance, overlaps with the temperament trait of negative emotionality. Conscientiousness, which reflects general self-control, is very similar to temperamental effortful control. The two remaining personality factors, agreeableness and openness to experience, do not have direct conceptual overlap with any temperament higher-order factor, but some of the characteristics of each of these personality factors can be found in lower-order temperament traits. For example, agreeableness, which reflects individual differences in the capacity to form and maintain positive relationships, is negatively related to anger reactivity and positively related to early attention (Caspi & Shiner, 2006). Much of the recent research on the links between temperament and personality has been conducted with the goal of demonstrating that personality can be measured in children (Halverson et al., 2003; Shiner & Masten, 2008). Studies on the structure of child personality have found personality in children to be structurally similarly to adult personality, increasing in similarity with age (Soto, John, Gosling, & Potter, 2011). Likewise, several studies have measured personality in children as young as preschool (e.g., Halverson et al., 2003) and suggest a significant overlap with temperament at the trait level of personality. Using principal components analysis on three temperament questionnaires and one personality instrument, De Pauw and colleagues (2009) recently demonstrated that many of the temperament dimensions

factored with the personality items to produce a six-factor model distinguishing sociability, activity, conscientiousness, disagreeableness, emotionality, and sensitivity. Interestingly, these new factors were more differentially predictive of internalizing and externalizing problem behaviors than the individual temperament and personality measures. Shiner and Caspi (2012) have argued that temperament traits develop into personality traits through four developmental processes that emerge at different times in childhood. Learning processes, which include positive and negative reinforcement, are proposed to begin early in life and be influenced by temperament traits. Likewise, environmental elicitation processes operate early in infancy and describe how the child's temperament elicits specific responses from others that may influence their behaviors over time. For example, infants who are high in positive emotionality elicit positive responses from parents, which may in turn affect their developing relationship. The third process, environmental construal, which is proposed to emerge after significant cognitive maturation, describes how temperament shapes the way children think about their environment and experiences. Fearful children, for instance, construe novel situations as stressful and overwhelming while fearless children interpret the same situation as exciting and stimulating. Environmental selection and manipulation, the fourth developmental process by which temperament traits influence personality, defines the ways in which temperament guides or shapes children's selection of which environments to engage. With age, children are given more opportunities to choose their environments, and it is proposed that temperament affects this choice. Thus, children who are more shy and withdrawn may be more cautious in their choice of friends. In sum, strong arguments have been put forward that temperament and personality are linked. However, empirical evidence is still needed to clarify the processes by which temperament influences or “is” personality. Although some studies indicate that temperament develops into the traits of personality, there is little data on how child temperament influences the other aspects of personality, such as self-concept and attitudes. This information will be important toward understanding whether temperament influences the development of psychopathology through its relation to personality (Clark, 2005).

The Measurement of Temperament The measurement of any construct should follow closely the theoretical assumptions of that construct. As temperament is defined broadly as individual differences in the intensive and temporal parameters of emotion, activity, and attention/regulation, temperament researchers must be able to assess the subtle variations of child behavior, subtleties that cannot be reported by the children themselves. Consequently, the measurement of temperament has relied almost entirely on questionnaires aimed at parents' ratings of their children's behavior, particularly in infancy and childhood. Contemporary temperament researchers rarely interview parents about their children's behavior, but such interviews had a critical role in the development of questionnaires in use today (Thomas et al., 1968). Another method for assessing temperament is to observe children either in a natural situation, such as their home or school, or in a laboratory setting where target behaviors are elicited. Finally, with technological advances in

electrophysiology, neural imaging, gene identification, and neuroendocrine assessment, researchers are now able to assess biological activity proposed by some theoretical models to underlie temperament. Next we describe each of these methods and discuss their advantages and disadvantages. We review the research on the neurobiology of temperament in a separate section later in this chapter.

Questionnaire-Based Assessments of Temperament Likely because of the ease of their administration, a number of questionnaires have been developed, most of which are based on the four contemporary models of temperament. These instruments differ, quite obviously, on the dimensions they assess but also on several other factors, including: (1) the number of items comprising the dimension of interest, (2) the specificity of the behavior and its context, (3) the reference group (e.g., comparison to other children), and (4) attention to developmental period and changes across development. One of the earliest parent-report temperament instruments developed were those based on Thomas and Chess's (1977) New York Longitudinal Study including the Infant Temperament Questionnaire (Carey, 1970), the Behavioral Style Questionnaire (McDevitt & Carey, 1978), and the Middle Childhood Questionnaire (Hegvik, McDevitt, & Carey, 1982). These questionnaires assess the nine dimensions proposed by Thomas and Chess as well as the broad factor of “difficultness.” Although used quite extensively across a number of studies, these instruments have demonstrated marginally acceptable psychometrics (Rothbart, Chew, & Gartstein, 2001; Vaughn, Bradley, Joffe, Seifer, & Barglow, 1987). Designed to assess temperament dimensions that are heritable, A. Buss and Plomin (1975) developed the EASI Questionnaire as well as the Colorado Childhood Temperament Inventory (Rowe & Plomin, 1977). Due to the low number of items (30–50), these measures have been popular with researchers interested in assessing temperament; however, many of the items are decontextualized (e.g., “child gets upset easily”), and the instruments do not take into account developmental changes in behaviors. Perhaps the most thoughtfully constructed and researched set of questionnaires are those developed by Rothbart and colleagues. Based on their psychobiological approach, these questionnaires are tailored to ask specific questions about the child's behavior within different contexts. On the Infant Behavior Questionnaire–Revised (IBQ-R; Gartstein & Rothbart, 2003), caregivers are asked questions about the child's emotional and motor reactions to, for example, being bathed within the last 2 weeks. Importantly, Rothbart has designed questionnaires for each developmental level. Aside from the IBQ-R, caregiver reports on toddlers (Early Childhood Behavior Questionnaire; Putnam et al., 2006), preschool/school-age children (Child Behavior Questionnaire; Rothbart, Ahadi, et al., 2001), and a self-report instrument in adolescence (Early Adolescent Temperament Questionnaire; Capaldi & Rothbart, 1992) have been tested and found to have good to excellent reliability and validity (Putnam et al., 2001). Factor analysis of each questionnaire has produced surprisingly consistent superfactors of surgency, negative affectivity, and orienting/effortful control with interesting shifts of some of the dimensions from one factor to another with development (Putnam et al., 2001). These changes reflect a more mature, organized system that is less reactive and more self-regulatory

over time. The advantages of questionnaire measures are numerous and include the ease of administration, the low cost in terms of time and money, and the ability to tap into infrequent behaviors that are descriptive of a child's temperament. And because parents observe their infants across a number of contexts, parent reports also provide a fuller, more ecologically valid description of their child's behavior than other methods. Indeed, although some instruments do not consider context in their questions, we can assume, for example, that parents are able to judge the likelihood that their child will respond negatively to novel persons because they view them across many different settings. Despite the advantages of temperament questionnaires, the accuracy of parent reports has been the topic of much debate (see Kagan & Fox, 2006; Rothbart & Bates, 2006). Parents may be biased in their view of their children, have differing interpretations of the questions posed, and have varying experience with other children for comparison purposes (Kagan, Snidman, McManis, Woodward, & Hardway, 2002). Studies on parental bias have convincingly demonstrated that many factors influence how parents rate their children's temperament, a fact that has led some researchers to advocate either disregarding parent report all together (Kagan, 1998; Seifer, Sameroff, Barrett, & Krafchuk, 1994) or using them in conjunction with other measures, such as laboratory observations (Rothbart & Bates, 2006; Stifter, Willoughby, & Towe-Goodman, 2008). The question of validity of parent report is largely due to the small to insignificant association between parent ratings and laboratory observations of child temperament. Although several studies have reported significant correspondence, the correlations are in the weak to modest range (rs < .30). Moreover, a similar number of studies have reported no such association (Mangelsdorf, Schoppe, & Buur, 2000). These findings have prompted many researchers to examine the source of disparity between what parents observe and what is observed in the laboratory. As mothers are typically the parent who reports on the child's behavior, the preponderance of these studies have focused on the characteristics of the mother. For example, maternal depression (Leerkes & Crockenberg, 2003; McGrath, Records, & Rice, 2008), anxiety (Vaughn et al., 1987; Voegtline & Stifter, 2010), and personality (Bates & Bayles, 1984; Hayden, Durbin, Klein, & Olino, 2010) have been related to mothers' reports of their children's temperament. Such factors may alter parent perceptions of their children's behavior, thus impacting their ability to objectively report on their children's emotional reactions and regulatory behavior. Indeed, in some studies, maternal characteristics were more strongly related to the mother's ratings of the child than with observed child behavior (e.g., Sameroff, Seifer, & Elias, 1982). Another reason for this disagreement between parents and observers is that the contexts in which they observe the child are vastly different. As noted, mothers have access to a variety of different settings while observers may see the child only in the laboratory or over a few hours in the home. In an attempt to control the conditions under which parents and observers rate a child's temperament, researchers compared parent and observer ratings under the same conditions. Seifer and colleagues (1994) had both parents and observers rate infant reactions

to the same tasks using the same rating system. The findings demonstrated that even after aggregating multiple observations, parents and observers did not agree. In a second study, Seifer (Seifer, Sameroff, Dickstein, Schiller, & Hayden, 2004) had mothers and observers rate the mother's own child and another child. Observers' and mothers' ratings of temperament showed strong convergence when rating another child but poor agreement when rating the mother's own child. Expanding on this methodology, Durbin and Wilson (2012) had mothers and naive observers rate children during the same tasks and compared both to objective coding of the behavior. Mothers and observers had weak agreement, and observers agreed with objective coding more than mothers. To explain this weak convergence, the researchers examined the effect of maternal mental health on mothers' ratings and found that mothers higher in depression and anxiety were more biased toward viewing their children more negatively than both naive observers and objective coders. Thus, even under conditions that are designed to maximize agreement, maternal characteristics operate to bias mothers' views of their children's temperament. Researchers should be aware of these biases when designing studies and consider using multiple informants and measures. We discuss this issue later in this chapter.

Observational Measures of Temperament A second method by which researchers assess the temperament of young children is through observation, in a laboratory setting or the more naturalistic settings of the child's home, daycare, or school. Typically, child behavior is elicited through a series of tasks designed to provoke a temperament dimension. To illustrate, children's approach/withdrawal tendencies and their accompanying emotional responses might be assessed by having an unfamiliar person interact with children or putting children in uncertain situations, such as the “risk room” (K. Buss & Goldsmith, 2000; Rothbart & Goldsmith, 1985), which contains structures/objects that may be unusual or “threatening” (large black box, tunnel). Variations in children's behavior, such as proximity to the black box, willingness go through the tunnel, and the number of spontaneous vocalizations exhibited during the interaction with the unfamiliar person, are then coded, usually from a video recording. There are many advantages to using observations to measure temperament. For one, observations that are conducted in the laboratory standardize the procedures, tasks, and setting across subjects, which cannot be expected from parent reports. That is, the contexts even with questionnaires such as the IBQ-R are not identical across families and thus introduce error into the assessment. A second advantage is that laboratory assessments target specific temperament behaviors that may not be as frequently observed by parents. Indeed, some parents may limit, for example, the number or type of novel situations their child is exposed to, which in turn can alter the expression of their fear reactivity. Laboratory measures, in contrast, may specifically create such situations to examine the range of reactions to novelty. A third advantage is that laboratory observations are usually recorded, which allows the researcher to examine the temporal and intensive qualities of behaviors that are related to temperament. Additionally, researchers can code different modalities of expression (facial, vocal) that may be less accessible to parents. It may be more difficult for a parent to recall the latency with which their

children become facially and vocally angry, for example. Despite the many advantages, observations are not the gold standard, as suggested by studies that examine the relationship between parent-rated questionnaires and laboratory measures. The lack of agreement might be due to problems in both measurement approaches or to differences in what is being measured by each approach (Goldsmith & Gagne, 2012). So, while data drawn from observations are less biased than that drawn from questionnaires, observational methods have their own issues. For one, observational measures do not capture temperament behaviors across a wide range of contexts. If we accept that temperament reflects cross-situational responses, then assessing behavior across more than one situation is essential. Observations provide a snapshot of children and thus may not be truly representative of their temperament. Second, the ecological validity of the laboratory setting is questionable. The unfamiliar laboratory and experimenters, no matter how friendly, can influence children's reactions to many of the tasks designed to evoke temperament, especially in more fearful, inhibited children. Unless one is interested in measuring behavioral inhibition, researchers must be aware of the degree to which the laboratory setting may affect behavior. Often researchers will have a warm-up period and allow children access to the parent at all times to alleviate this concern. A third disadvantage is the cost of conducting laboratory assessments of temperament, especially in terms of time. Parents must travel to the lab, observations must be video-recorded, and behaviors must be coded before data are available for analysis. Even if researchers decide to code online, which is not standard practice, the time it takes to train coders to become reliable on these coding schemes is substantial. Naturalistic observations are more ecologically valid, but they have similar disadvantages in that researchers are unfamiliar and the introduction of a person and/or camera in the setting may change aspects of the home or school. In addition, the setting is less controlled, which may affect the ability to administer standardized measures. Despite these disadvantages, observational measures of temperament have provided exceptional insights into the temporal and intensive features of individual differences in emotionality, activity, and regulation, particularly as it relates to behavioral adjustment. Many studies have employed observations to assess temperament, particularly if children's temperament is the primary construct. Because of the time and effort required to collect and code observational data, temperament researchers usually focus on one or two temperament dimensions. Kagan (1994), Fox et al. (2001), and others have used observational measures to examine behavioral inhibition, while Stifter (Dollar, 2011); Putnam & Stifter, 2005; Stifter, Putnam, & Jahromi, 2008), and others (Degnan et al., 2010; Pfeifer, Goldsmith, Davidson, & Rickman, 2002) have used tasks to elicit exuberance. Observational measures have also been used to measure anger reactivity/frustration (Braungart-Reiker & Stifter, 1996; Calkins, Dedmon, Gill, Lomax, & Johnson, 2002) and positive affect (Aksan & Kochanska, 2004; Goldsmith & Campos, 1990; Kochanska, Aksan, Penney, & Doobay, 2007). Effortful control has also been observed in a laboratory setting and presents an interesting contrast to measures of approach/withdrawal, emotionality, and activity level. As effortful control overlaps with the construct of executive function (Zhou, Chen, & Main, 2012), its assessment involves standardized tasks, such as the delay of gratification task or the Day/Night Stroop (see Other

Objective Measures of Temperament section). In an effort to provide standardized behavioral tasks that tap a list of temperament dimensions, Goldsmith and colleagues created the Lab-TAB (K. Buss & Goldsmith, 2000; Goldsmith & Rothbart, 1996; Goldsmith et al., 1993; Kertes et al., 2009) for premotor (infants), locomotor (toddlers), and preschool children. Not only are tasks provided for each dimension, but coding schemes for the parameters of latency, duration, and intensity of the behaviors are included. As an example, for assessing anger reactivity in toddlers, the Lab-TAB offers several frustrating tasks, such as the gentle arm restraint, confinement to a car seat, and a toy behind a barrier, which can then be coded from video recordings for the latency, duration, and intensity with which toddlers display the facial, vocal, and bodily expressions of anger. As the Lab-TAB is developmentally sensitive, the preschool version assesses anger reactivity with different tasks but uses the same parameters of emotional reactivity. Other dimensions that can be assessed using the Lab-TAB include positive affect, fear, sadness, shyness, approach, activity, persistence, and inhibitory control. Just as they have used questionnaires, researchers have been using observational measures to derive temperament dimensions or factors. Because of the difficulty and cost of conducting multidimensional longitudinal studies using observational methods, few studies have addressed the psychometrics of this approach. In one of the few existing studies examining the psychometrics of observed temperament measures, Durbin and colleagues (2007) examined positive and negative emotionality and their lower-order traits (e.g., sociability, interest, sadness, anger, fear) assessing behaviors from 12 tasks designed to elicit these traits at 3, 5, and 7 years of age. Results from this study indicated good to excellent stability as well as discriminant validity. Similar levels of short-term stability were reported in another study using Lab-TAB episodes to measure the dimensions of anger, fear, sadness, exuberance, and activity level (Majdandzić & van den Boom, 2007). Consistency across situations, as well as moderate convergence with parent ratings (both mother and father), was found for the observational measures. Using 12 episodes of the Lab-TAB during home visits, Gagne and colleagues (Gagne, Van Hulle, et al., 2011) created nine higher-level temperament composites mirroring those derived from questionnaire-based assessments. Cross-method convergence was found for most measures and parent-rated temperament, although postvisit ratings by observers were more closely related to the micro-coded measures. Predictive validity to a number of socioemotional outcomes has also been demonstrated by observational studies, many of which examined behavioral problems. These studies are reviewed later in this chapter. An alternative to the more microanalytic observational approach is to have experimenters observe behaviors across a laboratory or home visit and then globally rate specific aspects of the child's temperament (Braungart, Plomin, DeFries, & Fulker, 1992; Matheny, 1983; Saudino, 2009; Voegtline & Stifter, 2010). The majority of these measures are based on the Infant Behavior Record/Behavior Rating Scales, a behavioral assessment of the child while being administered the Bayley Scales of Motor and Mental Development (Bayley, 1969, 1993). Such items as social responsiveness, affect, activity level, irritability, reactivity, and task orientation are rated using a 9-point scale. Because these measures use global ratings and are based on

behaviors across tasks and during transitions between tasks, they are proposed to mimic parent ratings of temperament (Stifter et al., 2008). However, in one study, when compared to parent ratings, only measures of positivity, not negativity, converged (Stifter et al., 2008). In another study that compared observers' ratings with parent ratings and coded behaviors, observer ratings converged more strongly with coded behaviors than did parent ratings of child temperament (Gagne, Van Hulle, et al., 2011). That observer ratings across the lab/home visits were correlated with microcoding of behavior suggests that this method might be substituted for the more time-consuming coding approach to observing temperament. Although temperament observed in the laboratory/home and rated or coded by observers unrelated to the child may more closely reflect the true nature of the child than parent ratings, it does not necessarily present as the best alternative to parent ratings. Aside from being costly and limited in context and the types of behaviors that can be elicited, observational data may be influenced by other factors related to the raters/coders, such as experience with variations in child behavior and less sensitivity to the nuances of the temporal and intensive aspects of temperament. Training and inter-rater reliability calculated regularly can increase the accuracy of the ratings/coding systems. Likewise, reducing behaviors to observable parts that do not require significant judgment can also eliminate error. Researchers interested in observational studies should consider these limitations when designing studies on temperament and developmental psychopathology and when interpreting results from studies using this methodology.

Other Objective Measures of Temperament The most objective measure of a temperament dimension is one that does not use individuals close to the children to rate their behavior or observers to code behavior elicited in the laboratory, home, or school. One temperament dimension amenable to such a method is activity level. Using actigraphs, a device designed to detect frequency and intensity of motion within the range of normal human movement, researchers have examined the heritability of activity level as well as its relation to hyperactivity–attention-deficit disorder (Ilott, Saudino, Wood, & Asherson, 2010). The only requirement for actigraphs is that the more actigraphs used (one per limb) and the longer they are used, the more reliable the data (A. Wood, Rijsdijk, Saudino, Asherson, & Kuntsi, 2008). In a study of twin toddlers, actigraphy was related to parent ratings and observations of activity level, but the degree of agreement was different, with parents being less accurate than observers (rs = .25 versus .67), even when assessments were done within the same context (Saudino, 2009). In addition, the overlap in genetic effects between parent and mechanical measures of activity level was modest compared to the robust shared genetic effect between observers and the actigraph data. With advances in technology, automated methods have been developed and applied to facial expressions of emotion, which are central to the emotionality component of temperament. Using computer vision software that codes the action units of facial expressions of emotion automatically (Cohn & Kanade, 2007), researchers have successfully coded dynamic expressions of emotion in adults. More recently, automated facial coding has been applied to infants during a mother-infant interaction, and the data demonstrated concurrent validity with a

manual anatomically based facial coding system and construct validity with naive raters (Messinger, Mahoor, Chow, & Cohn, 2009). Importantly, these systems can measure the intensive quality of a facial expression of emotion (Messinger, Mattson, Mahoor, & Cohn, 2012), a temperament parameter identified in several temperament models (e.g., Rothbart's [Rothbart, Ahadi, et al., 2001] high- and low-intensity pleasure dimensions and Thomas and Chess's, 1977, intensity of mood dimension). Finally, standardized tests can also be used to tap certain temperament dimensions, specifically inhibitory control and the superfactor effortful control. Although this method involves an experimenter, the standardized administration and scoring (e.g., number correct; errors of commission/omission) decreases the subjectivity that is found in both parent ratings and observations. Standardized tests from other research domains, specifically the area of executive function, have been used by temperament researchers to measure effortful control. This is because the concept of effortful control overlaps significantly with several core components of executive function (Zhou et al., 2012). Effortful control is defined by Rothbart and Bates (2006) as the ability to inhibit a dominant response in favor of a subdominant response, to plan, and to detect errors. Executive function, like temperament, is an umbrella term that describes top-down goal-directed processes that operate to control attention, cognition, and behavior (Blair et al., 2005). Thus, both effortful control and executive function describe voluntary control in the face of challenging or competing demands. Unlike most measures of temperament, developmental researchers interested in executive function have developed a battery of tests that have been accepted by the research community as reliable and valid measures of these cognitive processes. Consequently, temperament researchers have adopted these measures, including the Day/Night Stroop, Go/No Go, and the Delay of Gratification task, to measure effortful control (Cipriano & Stifter, 2010; Kochanska & Knaack, 2003; Kochanska, Murray, & Harlan, 2000).

Multimethod Approach to Measuring Temperament As can be seen from the previous discussion, the two most utilized methods of assessing temperament, parent ratings obtained by questionnaires and laboratory/home observational measures, have their advantages and disadvantages. Although parent ratings are easy to administer, cost less, and provide an assessment of children's temperamental disposition across a wide range of contexts, studies suggest that parents are biased in their assessments, and researchers should be cautious about these biases when using parent reports of temperament in their studies. This advice is well heeded by researchers who are interested in the association between child temperament and problem behaviors since many studies rely on parent report of behavior problems as well. Despite this caveat, temperament questionnaires can be useful, depending on the focus of the study. If one is interested in mother-child interactions, how a mother views her child may be important toward understanding how she interacts with him or her. In a study of preterm infants (Voegtline & Stifter, 2010), mothers who were anxious rated their preterm infants as more negatively reactive than mothers of full-term infants, even though these infants were not observed to be more negative during home observations. Knowing this bias may improve our

understanding of how the transactions between mothers and their preterm infants may lead to poorer socioemotional outcomes. If due to time and money constraints researchers need to employ parent ratings of children's temperaments, it is advised that they obtain information from multiple sources, such as daycare providers/teachers, as they may be less biased and base their ratings on child behavior in different contexts than parents might (Bishop, Spence, & McDonald, 2003; Goldsmith, Rieser-Danner, & Briggs, 1991). In addition, using questionnaires that ask about specific behaviors within a reasonable time frame that are tied to a context and do not require parents to compare their child to other children of the same developmental age are highly recommended. Rothbart's infant, toddler, childhood and adolescent questionnaires attempt to reduce parent bias through these means (Putnam et al., 2001). Likewise, the decision to use observational measures should be driven by the theoretical approach and the questions researchers are attempting to address. Although costly, observation of children's reactions in tasks designed to elicit the behaviors of interest are likely to be more accurate than asking parents about specific parameters of behavior that characterize the temperament trait and to recall those situations that elicited these reactions. Observational measures have been criticized for the lack of validity (Rothbart & Bates, 2006); however, more recent reports indicate that observations of behaviors reflecting temperament dimensions, particularly if they aggregate across a number of tasks, are indeed valid (Goldsmith & Gagne, 2012). Ideally, research in temperament should consider multiple methods for assessing this construct. As Rothbart and Bates (2006) suggested, these measures represent components of variance in child temperament; each method contributes to a shared (parents and observers agree) and individual (that which is unique to the method) assessment of the child (Rowe & Kandel, 1997). A recent study on the quantitative genetics of temperament, specifically activity level, supports this conceptualization. Saudino (2009) found only modest overlap for the genetic effects of activity level measured mechanically and by parent report in the same situation. These findings suggest that different measures may be engaging different processes. Following this call for multiple methods, a handful of studies have combined laboratory/home observations and parent report of child temperament. Using principal components analysis, studies examining behavioral inhibition (Kertes et al., 2009) and social withdrawal (Kiel & Buss, 2011) have combined laboratory observations with questionnaire data. In one case, parent ratings, observer ratings, and coded behavior during parent-child interaction were combined to assess task persistence; the resulting composite measure correlated well with age, cognitive performance and behavior problems (Deater-Deckard, Petrill, Thompson, & DeThorne, 2005). Most studies have combined parent and child report (Lengua, 2006) and mother and father ratings (e.g., Leve, Kim, & Pears, 2005). Even home and laboratory observations (Dougherty, Klein, Durbin, Hayden, & Olino, 2010) have been combined to measure temperament. The availability of statistical techniques amenable to quantifying shared and independent variance across methods, such as structural equation modeling, should allow for a more reliable measure of child temperament and thus more confidence in understanding the role of temperament in adaptive and maladaptive behavior.

Summary A number of methods for assessing child temperament exist. The most widely used and affordable method is parent ratings of child temperament. Due to reporter bias, however, findings based on parent report alone should be interpreted with caution. Although observational methods avoid reporter bias, they also have their limitations. Automated measures and standardized tests appear to avoid many of the pitfalls of both parent ratings and observations of temperament, but they do not assess every temperament dimension. Perhaps with the advancement in digital and mobile technologies, innovations in the measurement of temperament will emerge. For now, researchers, guided by their research questions, should carefully weigh the advantages and disadvantages of each method and consider utilizing multiple measures of temperament. It is also important for researchers to keep in mind that these measures are tapping into different components of child temperament and are not interchangeable, as each method may be assessing different processes (Saudino & Wang, 2012).

Temperament and Developmental Psychopathology There is a long tradition linking temperament to psychopathology. Galen, in adopting the ancient Greek and Roman belief in the four humors (blood, yellow bile, black bile, phlegm), proposed that an imbalance of humors led to the temperaments of the sanguine, choleric, melancholic, and phlegmatic types. The idea that temperament influenced one's vulnerability to develop emotional and behavioral disorders carried through to the nineteenth-century psychoanalytic theories of Freud and Jung (Kagan, 1994). Importantly, contemporary models of temperament were instigated by a search for how child characteristics affected the development of child behavior problems. The primary aim of Thomas and Chess (1977) was to understand how the children they saw in their clinical practice with similar behavior problems came from both dysfunctional and “good” families prompting them to consider the child's contribution to his or her development of these problems. Many scholars have hypothesized the means through which temperament and psychopathology are linked. Rothbart, Posner, and Hershey (1995), in the first edition of Developmental Psychopathology (Cicchetti & Cohen, 1995), proposed several ways in which temperament might contribute to psychopathology, developmentally and with regard to treatment. Here we review seven of the hypothetical pathways from temperament that would most affect the development of behavioral/mental disorders in children. 1. Extremes in temperament may constitute a behavioral problem, such as extreme shyness or attention deficits. 2. Relatedly, extremes in temperament may predispose a child to develop a related disorder, such as when a highly fearful child develops general anxiety disorder or when the lack of attention regulation leads to conduct disorder (CD). 3. Temperament may affect others to act in ways that put the child at risk for developing problem behaviors. High levels of infant irritability, for example, may frustrate a

vulnerable parent, who in turn may treat a child harshly, leading to the development of conduct problems. 4. Another pathway may be through self-regulation, such that deficits in self-regulation would allow the expression of a temperamental extreme, as in the case of an easily frustrated child who develops ODD. 5. Alternatively, overregulation of attention or behavior may lead to clinical levels of inhibition or anxiety. 6. A temperamental characteristic may heighten a child's response to a given event, for example, when a child's high negativity heightens his or her stress levels, leading to depression. 7. Another possible pathway is that a certain temperamental disposition may protect a child from developing psychopathology, such as when some level of fear protects against CD or when positive affect protects against trauma. Thus, there are a variety of ways in which temperament contributes to developmental psychopathology, and recent research in this area supports many of these pathways. The means through which temperament influences the development of psychopathology proposed by Rothbart and colleagues (Rothbart & Posner, 2006; Rothbart et al., 1995) are aligned with the models linking adult personality to psychopathology (Krueger & Tackett, 2003), which were recently applied to the developmental literature (Tackett, 2006). These include the scar model, the pathoplasty/exacerbation model, the spectrum model, and the vulnerability/predisposition model. The scar model proposes that the development of psychopathology changes the individual's personality such that multiple occurrences of depression can increase neuroticism. According to the pathoplasty/exacerbation model, personality will affect the course, severity, and presentation of a disorder. In a review of research linking child personality to developmental psychopathology, Tackett (2006) concluded that empirical tests of these two models are rare, likely due to the methodological challenges of measuring personality before the emergence of a disorder and continuing to measure personality and psychopathology longitudinally. The remaining two models, however, have been supported by recent studies examining child temperament/personality and psychopathology. The spectrum model hypothesizes that personality traits and characteristics of psychopathology lie on a continuum, thus sharing a similar etiology. For example, attention-deficit/hyperactivity disorder (ADHD) may lie on the extreme end of inhibitory control/impulsivity. This model maps onto the first two pathways proposed by Rothbart (Rothbart et al., 1995) just reviewed. Studies that have investigated the spectrum model have found shared genetic links between such temperament traits as negative emotionality and effortful control with externalizing problems and CD (Tackett, Martel, & Kushner, 2012). The spectrum perspective is similar to that of Clark (2005), who proposed that the temperamental dimensions of negative affectivity, positive affectivity, and disinhibition are the shared factors that underlie the link between personality and psychopathology. This hierarchical model hypothesizes that these temperament

dimensions differentiate through biological and environmental developmental processes into personality traits and that, at their extremes, they put the individual at risk for psychopathology. Clark's conceptualization of the link between temperament and psychopathology also aligns with the precursor model, a more specific spectrum model, which assumes a developmental sequence in that temperament is evidenced before the onset of a disorder. The temperament trait of negative emotionality, for example, would be an early manifestation of internalizing disorder (Klein, Kotov, & Bufferd, 2011). The vulnerability/predisposition model incorporates many of the proposed temperamentdevelopmental psychopathology links of Rothbart (Rothbart et al., 1995). In this model, certain personality traits are proposed to place an individual at greater risk for developing certain clinical disorders, such as when impulsivity, a characteristic of low conscientiousness (and temperamental surgency), is related to antisocial behaviors. This model assumes a distinction between temperament/personality and psychopathological disorders and includes the potential for the interaction between temperament/personality and environmental factors to increase risk for psychopathology. The vulnerability model is analogous to the diathesis-stress model discussed earlier. One of the challenges in testing the vulnerability model is that the measures of temperament, personality, and psychopathology overlap. We consider this issue later in this section.

Internalizing and Externalizing Problem Behaviors The most prevalent outcomes examined by research on the link between temperament and developmental psychopathology are the broad dimensions of externalizing and internalizing problem behaviors. Based on Achenbach's (Achenbach, Edelbrock, & Howell, 1987) conceptualization and empirical work assessing childhood behavioral problems, externalizing and internalizing problem behaviors form a hierarchical model that accounts for the systematic covariation among behavioral and emotional symptoms commonly observed in childhood. Externalizing behaviors, so called because they represent outward behaviors that are uncontrolled, include aggressive and rule-breaking behaviors (Achenbach & Edelbrock, 1978). Several subscales of the externalizing dimension can be derived from the Child Behavior Checklist (CBCL; Achenbach & Edelbrock, 1983), one of the most commonly used instruments assessing childhood behavior problems. Importantly, the CBCL overlaps with constructs defined in the Diagnostic and Statistical Manual of Mental Disorders (Tackett, 2006). ODD is characterized by negativistic, noncompliant, and argumentative behavior and is considered an earlier and less intense form of CD. CD is primarily characterized by rule violations, such as aggression, destruction of property, and theft. The third subscale, ADHD, is not included in Achenbach's externalizing broad band but can be derived from items measuring inattention, impulsivity, and hyperactivity. Internalizing behaviors are those that are directed inward and subsume the disorders of anxiety and mood (depression). These problem behaviors are less studied in early childhood due to the inability of children to report on their internal states and the reliance on parent or teacher observations of behaviors reflecting anxiety and depression (Achenbach, McConaughy, & Howell, 1987; Edelbrock, Costello, Dulcan, Conover, & Kala, 1986; Kolko & Kazdin, 1993).

Several studies have examined internalizing behaviors in children younger than 5 years, but there is some question as to whether they constitute problematic behavior or just more extreme temperamental distress or inhibition (Goldsmith & Lemery, 2000). In addition, the reliance on parent report also raises questions as to parents' interpretation of child behavior, particularly given what is known about the accuracy of parent report. However, by middle childhood, selfreports of internalizing problems are relatively valid and reliable (Ebesutani, Bernstein, Chorpita, & Weisz, 2011; J. Goodman, Meltzer, & Bailey, 1998) but should be considered along with other informants' observations (Achenbach, 2006). Symptoms of anxiety include excessive worry and irritability whereas depressive symptoms can include sad mood, loss of interest, and sleep and eating disturbances. Perhaps because of their disruptive nature, externalizing behaviors are reported more often and earlier than internalizing behaviors. Research indicates that children as young as 3 years of age are at high risk of developing clinical levels of externalizing disorders if a constellation of noncompliance, overactivity, poor regulation, and aggression occurs frequently beyond the age-appropriate period, increases in intensity, and cuts across several contexts and domains. Moreover, children who exhibit externalizing behaviors early and persistently are more likely to develop antisocial behaviors in adolescence and adulthood (Odgers et al., 2008). A six-site, cross-national study of physical aggression in childhood, for example, demonstrated that boys with developmental trajectories of consistently high levels of physical aggression were associated with both violent and nonviolent delinquency in adolescence (Broidy et al., 2003). It is important to note that those behaviors that put the child at risk for the development of externalizing behaviors (e.g., overactivity, poor regulation) are temperament traits which is consistent with the spectrum model of developmental psychopathology that proposes extremes in temperament to be pathological. The developmental trajectories for internalizing disorders mirror that of externalizing disorders in children, although they are more transient. Children who exhibit high and increasing levels of internalizing behaviors are more likely to be anxious later in childhood or adolescence (Feng, Shaw, & Silk, 2008; Ollendick & King, 1994). Likewise, adolescent anxiety or depression is a risk factor for the same difficulties in early and even late adulthood (Coleman, Wadsworth, Croudace, & Jones, 2007; Glied & Pine, 2002; Pine, Cohen, Gurley, Brook, & Ma, 1998), particularly if it is persistent in adolescence. One source of internalizing disorders is the presence and rapid increase in externalizing problems. Externalizing and internalizing disorders co-occur at greater than chance levels (Angold, Costello, & Erkanli, 1999). When examining the developmental trajectories of both externalizing and internalizing behaviors across development, the data suggest that externalizing disorders may lead to increased internalizing problems in early childhood (Gilliom & Shaw, 2004) and anxiety and/or depression in adolescence (Copeland, Shanahan, Costello, & Angold, 2009; Loeber & Burke, 2011). Some have argued that this progression from externalizing to internalizing problems may be the result of the accumulation of negative interactions arising out of the child's dysregulation that makes him or her vulnerable to anxiety or depression (Wolff & Ollendick, 2006). Many of the studies investigating the developmental emergence and course of externalizing and

internalizing problem behaviors have considered child, parent, and demographic markers of risk. Child temperament has been identified as one of those factors. However, a problem unique to the examination of the link between temperament and developmental psychopathology is the method used to assess both constructs: parent ratings. Not only are both assessments based on the same reporter, but the instruments may be confounded. Measurement Confounding Identified as a significant issue in research on personality and psychopathology, the confounding of measures of temperament and symptoms of psychopathology in children takes on more methodological importance because of the reliance on parent report. As with child temperament measures, the assessment of problem behaviors in children, such as the broad bands of externalizing and internalizing disorders and the more specific subtypes of ODD, CD, ADHD, anxiety, and depression, is typically accomplished through parent ratings of the child's behavior. Consequently, many of the studies examining the link between child temperament and developmental psychopathology utilize the same informant, typically the mother, which can artificially inflate the relationship between the two constructs. To address this issue, the removal of overlapping items has been recommended, but this may decrease the validity and reliability of the measures. Lengua, West, and Sandler (1998) proposed two reasons for this overlap in measures, one theoretical and the other methodological. The spectrum model of the temperamentpsychopathology association discussed earlier reflects the potential for similarity in items measuring temperament and symptoms of developmental psychopathological disorders. As many of the externalizing and internalizing disorders constitute emotional and behavioral reactivity at the extreme, overlap among measures might be expected. The dispositional/vulnerability model proposes that temperament and behavior problems are distinct but related; thus, any overlap in measures would be an issue for testing this model as well (Lemery, Essex, & Smider, 2002). In addition, overlap might be expected because the development of each temperament and problem behavior measure was done independently without consideration of the items tapping the other construct. Consideration of overlap in items would increase discriminant validity (Lengua et al., 1998). At least three studies have directly examined the measurement confounding issue using similar techniques (Lemery et al., 2002; Lengua et al., 1998; Sanson, Prior, & Kyrios, 1990). Using both expert ratings and factor analysis, several items in the temperament measures and symptomology measures were confounded. In one study, 38% of the symptom items and 10% of the temperament items were rated by experts as confounded; factor analysis of caregiver reports revealed overlap in 23% of symptoms and up to 14% of temperament items (Lemery et al., 2002). After eliminating confounding items, each of these studies reanalyzed the correlation between the “purified” measures of temperament and symptoms and found reduced but still significant association, even when tested longitudinally (Lemery et al., 2002). Taken together, the findings from these studies (Lengua et al., 1998; Lemery et al., 2003; Sanson et al., 1990) suggest that the overlap in items measuring temperament and behavioral

problems does not affect their relationship. Many of the studies to be reviewed utilized parent reports of both child temperament and problem behaviors, concurrently and longitudinally. However, although simple correlations indicate some association, as predicted, the results should continue to be viewed with caution as the reporter of temperament and symptomology is typically the same, usually the mother. Current research has, and future research should, continue to consider multiple informants as well as observational measures.

Temperament and Psychopathology: Review of the Research Literature In this section, we discuss research that has addressed the link between various temperament traits and styles and the development of behavior problems and early psychopathology. Most research considering the association between temperament and different forms or precursors of psychopathology has measured direct, linear effects. In such studies, the research questions address how a single temperament trait contributes to the development of behavior problems or psychopathology. Additive effects are also possible in which multiple temperament traits linearly increase the child's risk of developing a problem behavior, such as when fearfulness or negative emotionality predict children's internalizing behavior problems (Gilliom & Shaw, 2004) There are also indirect processes by which temperament is associated with psychopathology. We discuss two forms of these indirect pathways: mediation and moderation. Research on mediated associations largely has focused on the manner in which temperament affects a child's interactions with the environment, which, in turn, shape the child's development of psychopathology. For example, Eisenberg and colleagues (Eisenberg, Zhou, et al., 2005) found support for the role of effortful control in mediating the relation between parental warmth/positive affect and children's later externalizing behavior problems. The second indirect process widely used in temperament studies involves investigating the conditions by which certain temperament traits are associated with behavior problems and psychopathology; specifically, these investigations consider moderating factors on the relation between temperament traits/styles and psychopathology. There are two central types of moderation that we consider in this review due to their large presence and influence: temperament × environment interactions and temperament × temperament interactions. It should be noted that studies on temperament × environment interactions hypothesize the moderating effects differently; some research investigates the moderating effects of the environment (i.e., parenting behaviors) and others measure the moderating effects of temperament. Although these views are complementary, we largely focus on studies investigating the moderating role of the environment on the relation between temperament and developmental psychopathology. It is important to note two fundamental issues when reviewing the research literature linking temperament to psychopathology. First, as previously discussed, it is critical to keep in mind that there is potential confounding of measurements of temperament and behavior problems in much of the existing literature (Lemery et al., 2002; Lengua et al., 1998). This is especially true

with research solely using parental reports of children's temperament and behavior problems. Thus, it is essential to take into account that the association between temperament and behavior problems may be overly inflated in some existing research. Another point to consider is the difficulty determining the direction of influence in a large majority of this research. More recent temperament studies have used statistical and methodological techniques to provide information regarding the direction of effects from temperament to psychopathology; however, this point warrants attention when reviewing the literature. Finally, it is critical that the reader consider the sample employed within each study. Overall, most research is best generalized to community-based samples rather than clinically referred populations since most studies include only a few children who meet the cutoffs for clinically significant behavior problems. Indeed, the reader should keep in mind that the number of investigations focused on clinically significant levels of psychopathology in children is quite small in most of the research reviewed. With that said, the basis of a developmental psychopathology perspective is to consider both normal and abnormal, adaptive and maladaptive developmental processes (Cicchetti, 1984, 1990, 1993; Cicchetti & Toth, 2009; Rutter, 1986; Sroufe, 1990), and, thus, the examination of the role of temperament in the development of psychopathology in both normative and at-risk populations is important. We organized our review of the research on the links between temperament traits/styles and the development of behavior problems and early psychopathology around the three broad temperament factors of negative emotionality, surgency, and effortful control (Rothbart & Bates, 2006). Within those broad categories, we also included research examining temperament types, specifically the exuberant and behavioral inhibited temperamental types, which have been extensively researched in the last 20 years. We also organized the review to include studies that focus on the specific temperament traits included in the superfactors that have also been theoretically and empirically related to developmental psychopathology. Thus, after reviewing the research on negative emotionality, we review the research on fear reactivity and anger reactivity. Likewise, we review the temperament traits of positive emotionality and activity level after reviewing the surgency research literature. Within each section, we first provide a broad description of the temperament factor/trait and then review the literature outlining direct linkages between the temperament trait and behavior problems/psychopathology. Next, we review relevant studies examining temperament × environment interactions in predicting behavior problems/psychopathology. Each section concludes with a summary and brief discussion of future directions. We end with an overview of temperament × temperament interactions that have been the topic of some of the most recent research in this area.

Negative Emotionality Negative emotionality is a temperament factor that includes the emotions of frustration/anger, fear, anticipatory anxiety, sadness, guilt, and discomfort. Children high in negative emotionality are prone to cry, fuss, and show more negative emotions overall, as well as display more intense emotional reactions in a variety of situations (Rothbart & Bates, 2006). Historically, research has linked temperamental negative emotionality to various aspects of later

development, including psychopathology (Rothbart & Bates, 2006). Indeed, this temperament factor has been shown repeatedly as an important predictor of both externalizing and internalizing behaviors across development, beginning in infancy (Bates, Bayles, Bennett, Ridge, & Brown, 1991; Eisenberg, Sadovsky, et al., 2005; Hagekull, 1994; McClowry et al., 1994; Rothbart, 2011; Rothbart & Bates, 2006). Although the direct relations between negative emotionality in the first years of life and internalizing and externalizing behaviors are modest, replication of these findings across various studies provides support that these are reliable associations. Garstein Gartstein, Putnam, and Rothbart )2012) found that children's externalizing problems in preschool were predicted by high negative emotionality in infancy, toddlerhood, and preschool, providing evidence for the influential role of negative emotionality across early development. Indication of the importance of negative emotionality across early childhood in predicting internalizing behavior problems has also been provided. In the Uppsala Longitudinal Study, for example, parent-reported negative emotionality when the children were 10 to 15 months old predicted children's self-reports of internalizing and externalizing behaviors when they were 4 and 8 to 9 years old (Bohlin & Hagekull, 2009). Although negative emotionality predicted both internalizing and externalizing behaviors in this study, there was a much stronger relationship between negative emotionality and internalizing behaviors. Of note, negative emotionality was not a significant direct predictor of anxiety or depression when these individuals were 21 years old, although there were other direct and indirect temperament effects. In addition, some research has examined negative emotionality as it relates to trajectories of behavior problems. This line of research is based on evidence that early behavioral and emotional difficulties often precede psychopathology later in development (e.g., Caspi, Moffitt, Newman, & Silva, 1998; Hofstra, Van der Ende, & Verhulst, 2002; Mesman & Koot, 2001) and are critical for our understanding of the stability, growth, and continuity of behavior problems across development. Gilliom and Shaw (2004) found in a study of disadvantaged boys that children's expression of negative emotionality predicted their levels of externalizing behaviors at the first time point (2 years of age), but it was not associated with children's changes in externalizing behaviors from age 2 to age 6. Importantly, this study employed multiple methods of observed and maternal report of children's negative emotionality, maternal report of behavior problems at age 2, and maternal and teacher report of behavior problems at age 6, which, as previously discussed, adds to the validity of this study. Findings from this study suggest that children's negative emotionality may contribute to the onset of their early externalizing behaviors but perhaps not its continuity. Even though there has been extensive research outlining the association between negative emotionality and later psychopathology, more recent research has acknowledged the importance of examining the specific negative emotions of anger, sadness, and fear that are commonly combined in measures of negative emotionality. Indeed, researchers have pointed to the differential roles of these specific negative emotions in children's developmental trajectories toward behavior problems and later psychopathology; thus, the importance of these specific negative emotions will be reviewed.

In addition to studies investigating the direct association between negative emotionality and psychopathology, researchers have also considered the interaction between negative emotionality and aspects of the child's environment in predicting children's development of behavior problems. Within this literature, there is a strong emphasis on parental personality and parenting behaviors as the measure of “environment.” For example, mother's internalizing affect moderated the relation between toddler's negative emotionality and depression concurrently, such that children high in negative emotionality who had mothers with higher levels of internalizing symptoms were rated as showing greater depressive symptoms than highly negative children whose mothers were not high on internalizing symptoms. Interestingly, this association was not found for children's anxiety (Marakovitz, Wagmiller, Mian, BriggsGowan, & Carter, 2011). Research has also indicated moderating effects on the relation between negative emotionality and externalizing behavior problems. Infant males high in negative emotionality, according to both parent report and laboratory observations, were more likely than boys low in negative emotionality to be rated as high in externalizing behaviors at age 3 if their parents also showed high intrusiveness and negative affect (Belsky, Hsieh, & Crnic, 1998). Importantly, results from this study incorporated both parent report and laboratory observations of children's negative emotionality. Additional research examining this topic area has been conducted with a similar, yet distinct, temperament construct, difficultness, which is reviewed in the next subsection. Examination of trajectories of externalizing behaviors, not just measures of externalizing behaviors at one point in time, has provided support for the interaction between negative emotionality and parenting in predicting behavior problems, as well. Owens and Shaw (2003) reported that children low in negative emotionality at 18 months of age with mothers high in depression had a faster decline in externalizing behaviors from age 2 to age 6 than children high in negative emotionality. This finding suggests that low negative emotionality could be a protective factor when children are at risk for developing externalizing behaviors due to their mothers' high depressive symptoms. Also of interest, results from this study showed that when mothers were low in depressive symptoms, the opposite finding was revealed. Children high in negative emotionality who had mothers low in depressive symptoms showed greater improvement in externalizing behavior over time so that by the time the children were 6 years old, highly negative children were comparable in externalizing behaviors to children low in negative emotionality. Drawing from the same study but examining different maternal behaviors, Gilliom and Shaw (2004) showed that the combination of negative emotionality, fearlessness, and negative maternal control was related to a high, nondecreasing trajectory of externalizing behaviors from age 2 to age 6. In general, this literature supports the widespread perspective that environmental risk, such as negative parenting behaviors, exacerbates the potentially deleterious effects of negative emotionality in the development of behavior problems (Bates & Pettit, 2007). However, the potentially maladaptive outcomes that are associated with negative emotionality can be tempered by positive, supportive, and warm parenting behaviors. It is noteworthy that a recent meta-analysis (Paulussen-Hoogeboom, Stams, Hermanns, & Peetsma, 2007) suggested that the relations between negative emotionality and parenting characteristics/behaviors in predicting

children's behavior problems are dependent on other factors, such as context (high-risk and low-risk samples) and the methodology employed in the study (e.g., micro- or macro-level approaches, modeling trajectories of behavior) and thus should be considered in future research. Further, as argued previously, the measurement employed in studies of temperament should be taken into account. Specifically related to the negative emotionality × parenting literature, Paulussen-Hoogeboom and colleagues (2007) found in their meta-analysis that forms of negative parenting behaviors were associated with children's negative emotionality only when parental reports were employed. This finding again highlights the importance of multiple methods and informants, such as using observational measures in conjunction with parent reports of child behavior. “Difficult” Temperament Temperamental “difficulty” is a construct that is closely related to negative emotionality. As previously discussed, Thomas and Chess (1977; Thomas et al., 1968) identified a group of children, labeled “difficult,” who were defined as high in withdrawal, negative mood, unadaptable to change, intense, and irregular (Thomas, Chess, Birch, Hertzig, & Korn, 1963). Historically, difficultness has received considerable attention by researchers interested in the relationship between temperament and maladaptive developmental outcomes, and the results support this association concurrently and longitudinally (e.g., Bates et al., 1991; Keenan, Shaw, Delliquadri, Giovannelli, & Walsh, 1998). However, this construct has been widely criticized (e.g., Plomin, 1982; Rothbart, 1982) and more recently put aside in temperament research in favor of more reliable measures of negative emotionality or more discrete temperament traits. One central issue with the use of the “difficult” label is that it identifies children as a problem early in life. The ramifications of attaching this negative characterization on children so early may cause unnecessary issues for them. For example, parental expectations that a child is difficult may influence their behavior toward the child, thereby worsening the child's behavior while reinforcing parental perceptions. This label may also influence the child's developmental trajectories in many ways, such that the difficult label may restrict the contexts the child is exposed to, thereby potentially affecting his or her experiences and opportunities. Further, the difficult label may also be inaccurate, as “difficult” behavior may cause problems in some contexts but be beneficial in others. Yet such behavior is less likely to be viewed as beneficial because of the negative connotation associated with the term difficult. Another issue with the use of the “difficult” label to describe child temperament is the inconsistencies in its operationalizations. In particular, there have been problems replicating the difficult temperament cluster, which has led to various forms of this construct. For example, algorithm-based perspectives (Fullard, McDevitt, & Carey, 1984; McDevitt & Carey, 1978) include rhythmicity, whereas Bates (1980, 1989) does not include rhythmicity and instead argues that the central aspect of temperamental difficulty is negative emotionality (Rothbart, 2004). Because of this issue, comparisons across studies can be problematic, and caution should be used when interpreting results from studies that examine “difficult” temperament since studies may be tapping into different aspects of a child's temperament (Rothbart, 2012).

Although there are issues with the use of the difficult temperament construct, a wide body of literature has examined its relationship with problem behavior. Because of this fact, we provide a brief review of some prominent studies that have measured a form of “difficult” temperament as well as more recent studies that have incorporated the differential susceptibility perspective using this construct. As mentioned, children labeled “difficult” are more likely to develop a variety of maladaptive developmental outcomes. For example, temperamental difficultness, measured by the Infant Characteristics Questionnaire as frequent and intense negative affect and attention demanding, has been shown repeatedly to predict children's later internalizing and externalizing behaviors (Bates, Maslin, & Frankel, 1985; Bates et al., 1991). Indeed, difficult temperament, as rated by the mother when the infant was 6 months of age, was predictive of conduct problems at age 3 (Bates et al., 1985), and a similar relation was found for report of children's externalizing behaviors at ages 7 to 8 (Bates et al., 1991). Further, infants rated as difficult and fussy at 6 months were rated as showing high levels of hyperactivity and aggression at age 17 (Olson, Bates, Sandy, & Lanthier, 2000). Other studies have found that temperamental difficultness, as measured by the Infant Characteristics Questionnaire, was predominantly associated with internalizing behavior problems, not externalizing behaviors (e.g., Keenan et al., 1998). As mentioned, many of these studies should be interpreted with caution, not only because they examined difficult temperament but also because both temperamental difficultness and behavior problems were both measured via maternal report. Similar to the negative emotionality × environment interactions, many studies have examined the interactive effects of difficult temperament and parenting and have shown that children with difficult temperaments are much more likely to develop behavior problems when their parents behave in a punitive, negative manner (Bates & Pettit, 2007; Kochanska & Kim, 2013; Rothbart & Bates, 2006). Recently the model of differential susceptibility has been utilized to provide greater understanding of the relation between difficult temperament and parenting in predicting children's developmental outcomes. Several studies support the hypothesis that difficult children are more susceptible to a range of parenting qualities, both positive and negative, than are easy children (Belsky, 1997; Belsky, Bakermans-Kranenburg, & van IJzendoorn, 2007; van IJzendoorn & Bakermans-Kranenburg, 2012b). For example, difficult children have been found to be more susceptible to negative parental discipline by showing more externalizing behavior problems as well as more susceptible to parental positive discipline by showing less physical aggression and fewer externalizing behavior problems than children with easy temperaments (van Zeijl et al., 2007). Additional support shows that children with difficult temperaments whose mothers were high in maternal sensitivity had the strongest decrease of externalizing problems from ages 2 to 5 (Mesman et al., 2009). More recently, Kochanska and Kim (2013) found that children's observed difficult temperament, as measured by poor effortful control and high anger proneness, moderated the relation between mothers' responsiveness at 30 to 33 months and children's externalizing behaviors at 40 months of age. For difficult toddlers, maternal unresponsive behaviors was related to higher levels of children's externalizing behaviors, whereas high levels of maternal responsive behavior with difficult children was related to low levels of externalizing behaviors. It is worth noting that the results from this study did not find that difficult children who received responsive parenting

were statistically lower in externalizing behaviors than were the easy children, thereby providing support for a diathesis-stress model, not differential susceptibility. Anger Reactivity It has been suggested by many that anger, sadness, and fear reactivity, encompassed in negative emotionality, might better be treated separately. Specifically, high levels of anger/frustration are hypothesized to underlie externalizing behaviors whereas greater sadness and fear reactivity are consistent with symptoms of internalizing problems (Rothbart & Bates, 2006). For example, in a sample of preadolescents, higher fear reactivity was associated with internalizing problems and higher anger reactivity was related to externalizing problems (Oldehinkel, Hartman, De Winter, Veenstra, & Ormel, 2004). It is important to note that this association is not perfect, as outlined next (i.e., anger sometimes is associated with internalizing behaviors). The association between anger and externalizing and aggressive behaviors has been found repeatedly across decades of research, providing considerable support for their relation (Eisenberg et al., 2009; Gilliom, Shaw, Beck, Schonberg, & Lukon, 2002; Lemery et al., 2002; Lengua, 2006; Rothbart, Ahadi, et al., 1994; Rothbart, Derryberry, & Hershey, 2000; Rydell, Berlin, & Bohlin, 2003). For example, infant frustration observed in the laboratory was related to parent-rated aggression when children were 7 years old (Rothbart, Derryberry, et al., 2000), and 5-year-olds high in parent-rated anger were more likely to be rated as high in externalizing behavior problems by teachers in preschool and elementary school and by parents at home (Rydell et al., 2003). Unfortunately, it is still unclear how the relation between anger and externalizing behaviors emerge developmentally. In particular, it is unknown if anger motivates externalizing types of behavior and/or if children who exhibit aggressive behavior become more angry and volatile across development because their aggressive behavior leads to peer rejection (Rubin, Bukowski, & Parker, 2006). Indeed, even though a functionalist view of emotions proposes that anger motivates goal-oriented behavior and therefore can serve an adaptive purpose (Saarni, Mumme, & Campos, 1998), inappropriate levels and/or expression of anger may generate aggressive behaviors that negatively affect peer interactions and prevent socially adaptive problem-solving abilities. Some research finds that children prone to high levels of anger early in development are at a higher risk of developing externalizing and internalizing problems later in childhood and adolescence than children low in anger (Eisenberg et al., 2009; Frick & Morris, 2004; Gartstein et al., 2012; Rothbart & Bates, 2006). However, these associations tend to be stronger for externalizing behaviors than for internalizing behaviors (Eisenberg et al., 2001; Oldehinkel et al., 2004). Although the link between anger reactivity and internalizing behavior is somewhat counterintuitive, the approach/withdrawal model of emotions proposed by Carver and Scheier (1998) suggests that sadness and depression can result from an individual repeatedly failing to approach a goal. Thus, children might initially experience anger when progress toward a goal is blocked, but after continued efforts to obtain this goal are hindered and presumed lost, the internalizing emotion of sadness may develop. An additional explanation for this association comes from research indicating that increases in externalizing

behavior problems coincide with increases in internalizing behavior problems across the preschool period (Gilliom & Shaw, 2004). Thus, there may be multiple pathways from anger reactivity to the development of both internalizing and externalizing behaviors. As outlined by Patterson (1976, 1982) in his coercion theory, interactions between parents and children that are negative and/or coercive in style are likely to persist and accumulate over time, thereby increasing children's likelihood of developing externalizing, aggressive behaviors. This could be especially true for children prone toward experiencing negative emotions, specifically anger. In addition, within hostile, harsh environments, it is likely that children prone to experiencing and expressing anger will not learn appropriate methods of regulating their emotions, thereby exacerbating their risk of developing behavior problems and later psychopathology. There is much correlational support for the notion that children's anger is negatively associated with supportive parenting. For example, anger is concurrently associated with lower levels of parental warmth in infancy (Kochanska, Friesenborg, Lange, & Martel, 2004) and increased levels of harsh parenting (Rhoades et al., 2011). Further, it is hypothesized that for anger-prone children, the effects of negative parenting behaviors, such as hostility, rejection, inconsistent discipline, and the like, might engender greater anger and aggressive behavior, characteristics of both internalizing and externalizing behavior problems. Thus, the effects of poor environmental conditions, such as negative parenting behaviors, might be especially detrimental to children temperamentally high in anger, given that low-quality and negative parenting seem to amplify behaviors already at risk (Calkins & Fox, 2002; Morris et al., 2002). In support of this position, Morris and colleagues (2002) found that first- and secondgrade children high in temperamental anger were rated as higher in teacher-reported externalizing behavior problems if children viewed their mothers as high in overt hostility. Interestingly, anger-prone children were also high in teacher-rated internalizing behavior problems when these children rated their mothers as high in intrusive control. Lengua (2008) reported similar results when examining change in problem behaviors. Children high in frustration had greater increases in externalizing behaviors across 1 year during middle childhood (average age at Time 1 was 8.5 years old) within the context of child-reported maternal rejection. Interestingly, inconsistent maternal discipline predicted increases in internalizing behaviors for children high in frustration. Person-centered methodological approaches also have been employed in research linking anger and the development of behavior problems. For example, Degnan, Calkins, Keane, and Hill-Soderlund (2008) created longitudinal profiles from age 2 to age 5 of children's disruptive behavior, which can lead to antisocial behaviors later in life (Broidy et al., 2003). The findings showed that children high in observed anger were more likely to be members of the high disruptive profile (high levels of disruptive behavior across time) when they had mothers who were high in control (adult-oriented goals and strictness/punitiveness). However, results from this study provided support that high physiological regulation, as indexed by vagal withdrawal, served as a protective factor for anger-prone children by leading toward a greater likelihood of being a member of the normative profile than the high disruptive profile. Thus, similar to findings on negative emotionality × parenting interactions, parenting behaviors high

in control and hostility seem to intensify the relation between children's anger and later behavior problems. It is important to note that the developmental process by which parenting moderates children's anger reactivity in predicting behavior problems is still unknown. It could be that children temperamentally prone to anger/frustration have difficulty internalizing rules of conduct, which elicits harsh parenting, increasing the likelihood of behavior problems. Or it is possible that children high in anger might respond with increased frustration and irritability to negative parenting behaviors, leading to a greater likelihood of developing behavior problems. The developmental transactional process for anger reactive children warrants additional research. Fear Reactivity Additional support for the importance of considering the specific components (fear, anger, sadness) of negative emotionality comes from researchers linking individual differences in fear, measured as early as infancy, and the development of internalizing behavior problems (Biederman, Rosenbaum, Bolduc-Murphy, & Faraone, 1993; Biederman et al., 2001; ChronisTuscano et al., 2009; Hirshfeld-Becker et al., 2008). For example, Gartstein and colleagues (2012) reported that fear in toddlerhood and childhood, as reported by mothers, predicted children's internalizing scores both cross-sectionally and longitudinally. In another study (Rydell et al., 2003), parent-rated temperamental fear was positively associated with internalizing behavior problems longitudinally, although it is important to note that both were parent-rated. Further, children's low regulation of fear was related to higher internalizing behavior problems, both at home when the children were 6.5 years old (parent-reported) and in elementary school when the children were 8 years old (teacher-reported). A more recent perspective on the relation between temperamental fear and the development of internalizing behavior problems proposes the importance of examining context. K. Buss (2011) found that 2-year-old children's observed expression of fear in a highly threatening context was not associated with anxiety symptoms when the children were in kindergarten. Instead, dysregulated fear, characterized by high fear across all contexts including low-threat contexts, was related to higher maternal-reported anxiety when children were in preschool as well as social withdrawal when they were in kindergarten. Importantly, although children's inhibition (i.e., shyness, withdrawal) was related to maternal reports of social withdrawal during kindergarten, dysregulated fear predicted variance in social withdrawal over and above inhibition. As an extension of these findings, K. Buss and colleagues (2013) found that children with greater dysregulated fear at age 2 consistently predicted social wariness and social anxiety disorder symptoms throughout the child's kindergarten year. Although predominantly discussed as a strong predictor of internalizing behavior problems, temperamental fear is also reported as negatively associated with externalizing problems and is hypothesized to protect children from developing aggressive behaviors (Rothbart & Bates, 2006). For example, infants' expression of fear in the laboratory has been shown to predict lower maternal reports of aggression when the children were 7 years of age (Rothbart, Ahadi, et al., 2000). One explanatory mechanism through which fear reactivity early in life is believed to counteract the development of externalizing behavior problems is through the development

of conscience, which is negatively related to children's externalizing and aggressive behaviors (Eisenberg, Fabes, & Spinrad, 2006; Kerr, Lopez, Olson, & Sameroff, 2004). Indeed, children high in temperamental fear have been found to demonstrate better internalization of rules and standards (Kochanska, 1995; Kochanska et al., 2001) as well as show more empathy/guilt (Kochanska, Gross, Lin, & Nicholas, 2002; Rothbart, Ahadi, et al., 1994). It is worth mentioning that although temperamental traits, such as fear reactivity, are believed to contribute substantially in the development of conscience, which may lower the risk of externalizing behaviors, there are multiple factors involved in this development. One in particular, parenting behaviors, is discussed next as a moderator of the relation between fear reactivity and children's conscience. Various parental behaviors have been examined for their moderating role on the relation between temperamental fear and the development of internalizing behavior problems. It has been suggested that fearful temperament encourages warm, supportive parenting behaviors (Lengua, 2006); however, there is also evidence that temperamental fear is not associated with either more or less parental support (Paulussen-Hoogeboom et al., 2007). Although it is not clear if infant and childhood fear reactivity is more likely to elicit a certain type of parenting, there is evidence that negative parenting behaviors increases the relation between temperamental fear and internalizing behavior problems. Gilliom and Shaw (2004) found that children high in observed and parent-rated negative emotionality and fearfulness with mothers high in observed negative control were more likely to develop increasing internalizing behavior problems from the ages of 2 to 6. Moderated mediation analyses have also been conducted in investigations regarding the association between temperamental fear and later anxiety. Kiel and Buss (2010) found that maternal protectiveness mediated the relation between fearful temperament and later anxiety, but only for mothers who were able to accurately predict when their child would be fearful. Another central line of research on temperamental fear has been conducted by Kochanska (1991, 1995, 1997; Kochanska, Aksan, & Joy, 2007). In this research, she looked at the interaction between parenting behaviors and children's fear and fearless behaviors in predicting the development of moral and behavioral self-regulation, an important component of behavioral adjustment and predictor of fewer problem behaviors (Calkins & Keane, 2004; Keenan & Shaw, 2003). Kochanska (1991, 1995) hypothesized that since fearful children have a low threshold for fear arousal, parenting behaviors that are gentle, positive, and cooperative in nature will promote children's internalization of rules and morals in fearful children. In support of this hypothesis, she found that when mothers of temperamentally fearful children showed higher levels of gentle, low-power discipline and guidance, their children displayed more committed compliance and internalization than fearful children of mothers who did not show these behaviors. Therefore, it was concluded that an optimal level of anxiety arousal is elicited in fearful children when mothers use gentle discipline, thereby promoting compliance and internalization (Kochanska, 1997). It is important to note that temperamental fear is proposed to underlie a number of behaviors that are included in temperament models, such as withdrawal, behavioral inhibition, and shyness, all of which have been studied in relation to the development of internalizing

behaviors, specifically anxiety and social withdrawal. Fear reactivity, for example, is a major component, along with other behavioral and physiological traits, of the behavioral inhibited temperamental type or category, which we consider next. Behavioral Inhibition As mentioned, the construct of behavioral inhibition is closely related to other temperamental traits, such as negative emotionality, fear, and shyness. In general, behavioral inhibition reflects the inability to appropriately handle uncertainty rather than a predisposition to fear (Kagan & Snidman, 2004). In the 1980s, Kagan and colleagues (e.g., García Coll, Kagan, & Reznick, 1984; Kagan, Reznick, & Snidman, 1987b) introduced the construct of behavioral inhibition to the developmental community. Since then a number of studies have examined its development, physiological bases, and relation to developmental psychopathology. Taking an extreme groups/categorical approach, Kagan defined behaviorally inhibited children as those who tend to withdraw and show negative affect in response to novelty. Research has shown that when presented with unfamiliarity, children who are behaviorally inhibited display high levels of negative affect, distress, wariness, and anxiety (Calkins, Fox, & Marshall, 1996; García Coll et al., 1984; Kagan, 1997) and are likely to exhibit social withdrawal or reticence in early childhood (Fox et al., 2001; Rubin, Burgess, & Hastings, 2002). Individual differences in emotional and behavioral reactivity to novelty, according to Kagan (Kagan et al., 1987b), are functions of excitability in the amygdala and proximate areas of the brain. This excitability is thought to generate heightened fear responses to unfamiliar and novel situations, people, and things. For example, adults who had been identified as inhibited in toddlerhood, compared to those individuals who had been identified as uninhibited, had a greater amygdala response to novel versus familiar faces (Schwartz, Wright, Shin, Kagan, & Rauch, 2003). Behaviorally inhibited children also evidence greater autonomic reactivity than uninhibited children (PerezEdgar & Fox, 2005a; Schmidt, Fox, Schulkin, & Gold, 1999). The results of multiple longitudinal studies have demonstrated moderate stability of behavioral inhibition across childhood (Kagan, Reznick, Snidman, Gibbons, & Johnson, 1988; Pfeifer et al., 2002; RimmKaufman & Kagan, 2005). Approximately 15% to 20% of children are likely to show high levels of behavioral inhibition (Fox, Henderson, Marshall, Nichols, & Ghera, 2005), and about half of these children continue to show indicators of behavioral inhibition across childhood (Degnan & Fox, 2007). Understanding the developmental course of children categorized as behaviorally inhibited in early childhood has been important to developmental psychopathology. In general, research has found behavioral inhibition to predict later social behaviors and behavior problems (Kagan, 1994; Kagan, Reznick, & Snidman, 1987a; Rubin et al., 2002; Schwartz, Snidman, & Kagan, 1999). But more specifically, children who were consistently labeled as inhibited developed social anxiety disorders (e.g., Biederman et al., 1990, 2001; Hirshfeld et al., 1992; Schwartz et al., 1999). Interestingly, children identified as having anxiety disorders as early as 3 years of age displayed higher levels of behavioral inhibition as well as lower levels of positive affect (Dougherty et al., 2013). As with other studies on the moderating role of the environment on the relation between

temperament and psychopathology, research focused on behavioral inhibition has examined the role of parenting behaviors in the development of psychopathology. Overall, this literature suggests that intrusive, controlling, and/or oversolicitous parent behaviors increase inhibited toddlers' risk of developing social reticence and children's wary, anxious behavior in social situations (Rubin, Cheah, & Fox, 2001; Rubin, Hastings, Stewart, Henderson, & Chen, 1997; Rubin et al., 2002; Van Leeuwen, Mervielde, Braet, & Bosmans, 2004), whereas maternal warmth, sensitivity, and responsiveness are related to lower levels of behavioral inhibition and more socially acceptable behavior (Hane, Cheah, Rubin, & Fox, 2008; Park, Belsky, Putnam, & Crnic, 1997). Conversely, there are also findings from studies on behaviorally inhibited children showing that sensitive, overly responsive parenting behaviors, sometimes discussed as overly solicitous, maintain children's behavioral inhibition. It may be that by catering to the child's proclivity to avoid unfamiliarity, over-solicitous parents may be fostering the continuity of inhibited behaviors (Kagan, Arcus, & Snidman, 1993; Putnam, Sanson, & Rothbart, 2002). Moreover, oversolicitous parents may be reinforcing children's behavioral inhibition by rewarding their children's initial signs of distress with parental warmth and precluding any opportunity for children to develop and use self-regulatory skills (Fox et al., 2005). Although the research is mixed depending on how the parenting behaviors are defined (i.e., sensitive versus intrusive), the findings indicate that parenting behaviors aimed at encouraging children to engage socially may protect behaviorally inhibited children from developing social anxiety by lowering their attention bias to threat and promoting exploration and engagement with other children (Fox et al., 2005; J. Wood, McLeod, Sigman, Hwang, & Chu, 2003). In turn, this mastery of the environment will assist in the development of self-regulatory behaviors. Similarly, Kagan (1994) has posited that if fearful children are forced to face and deal with situations that cause them discomfort (e.g., fear or frustration) within the context of responsive parenting, they will be more likely to learn to regulate these emotions and therefore show fewer fearful behaviors with time. It is worth noting that a few studies have examined the moderating effect of other nonparenting environmental influences. For example, Essex Armstrong, Burk, Goldsmith, and Boyce (2011) examined the effect of teacher-child relationships in the first grade on mental health outcomes at grade 7 for behaviorally inhibited children. Behaviorally inhibited children developed worse mental health symptoms under conditions of high teacher-child conflict but the lowest levels of symptoms when there was low teacher-child conflict. Therefore, this study provided support for the differential susceptibility hypothesis for inhibited children. Another nonparental environmental context that is associated with behaviorally inhibited children's development of anxiety is that of the child care context (e.g., Fox et al., 2001). In this study, highly negative infants at 4 months of age were less likely to become inhibited later in development when they were placed in nonparental child care for 10 or more hours per week. As stability of behavioral inhibition is a risk factor for internalizing behaviors, the effects of child care in lessening this risk is achieved by exposing behaviorally inhibited children to other children and adults, encouraging social interaction, and requiring the use of regulatory strategies. However, it is important to consider that these effects may differ depending on the level of positivity and supportiveness within the child care environment (Clarke-Stewart & Allhusen, 2002; Gazelle, 2006). For example, Morrissey (2009) reported that preschoolers who

experienced a greater number of child care environments, such as in-home care, center care, and care by grandparents, were rated by their parents as higher in internalizing behavior problems. Thus, the amount of child stress created by nonparental care should also be considered when examining the effects of child care on children varying on their levels of behavioral inhibition. Summary In sum, negative emotionality, a temperament factor that includes the emotions of frustration/anger, fear, anticipatory anxiety, sadness, guilt, and discomfort, has been shown repeatedly to be a reliable predictor of both externalizing and internalizing behaviors across development. Direct links between negative emotionality and behavior problems at one time point exist as well as the association between negative emotionality and trajectories of behavior problems. Although there has been extensive research outlining the association between negative emotionality and later psychopathology, recent work acknowledges the importance of examining the lower-order traits that make up the negative emotionality factor; indeed, evidence of the differential roles of these specific negative emotions in children's developmental trajectories toward behavior problems and later psychopathology has emerged. For example, there is considerable support for the association between anger reactivity and externalizing behavior problems. Although temperamental anger is more strongly related to externalizing behaviors than to internalizing behaviors, growing empirical evidence indicates that children prone to high levels of anger early in development are at a greater risk of developing both internalizing and externalizing problems later in childhood and adolescence than are children low in anger reactivity. Temperamental fear has been shown to be a good predictor of internalizing behavior problems. However, it is also important to note that fear is negatively associated to externalizing problems, which suggests that it may operate as a protective factor for aggressive behaviors. Behavioral inhibition, a temperamental type characterized by high levels of fear/uncertainty along with other behavioral and physiological traits, has also been linked to internalizing behaviors, such as social withdrawal and reticence as well as anxiety. The moderating role of the environment on the association between negative emotionality and later behavior problems has also been investigated. The findings support the widespread perspective that negative parenting behaviors exacerbates the maladaptive effects of negative emotionality in the development of behavior problems, whereas positive parenting behaviors can temper this relationship. The effects of parenting behaviors are thought to be especially detrimental to children temperamentally high in anger within the context of low-quality and negative parenting. These findings support the transactional nature of the parent-child relationship such that highly negative children are likely to evoke more negative responses from their parents, which may develop into a pattern of maladaptive interactions that decrease the children's opportunities for meeting developmental milestones, such as learning appropriate regulatory behaviors. Fewer studies have examined parenting and temperamental fear in predicting behavior problems. Parenting behaviors that are intrusive, controlling, and/or oversolicitous are related to temperamental fear, behavioral inhibition, and social

reticence, whereas maternal warmth, sensitivity, and responsiveness are related to lower levels of behavioral inhibition and more socially acceptable behavior. However, there is also research supporting the notion that sensitive, overly responsive parenting behaviors maintains children's behavioral inhibition, a risk factor for internalizing, by catering to children's proclivity to avoid unfamiliarity. Important research questions remain regarding the specific types of parenting behaviors that lower fearful children's risk of developing internalizing behavior problems. Overall, researchers have begun to address important questions related to the trajectories of negative emotionality as well as models of differential susceptibility, both of which are important areas of future research. Although the role of negative emotionality in the development of psychopathology is well established, we encourage future research to focus on individual differences in the expression and experience of the more specific emotions of fear, anger, and sadness rather than the broad constructs of negative emotionality and difficult temperament. In addition, despite these consistent and promising findings, it is still unclear how the relation between anger reactivity and externalizing behaviors and fear reactivity and internalizing behaviors emerges developmentally. For example, some studies suggest that anger and aggression share similar neural circuits (Coccaro, McCloskey, Fitzgerald, & Phan, 2007), adding to the complexity of our understanding of the association between anger and the development of behavior problems and psychopathology. Moreover, the existing research literature largely does not acknowledge the direction of effects. Thus, we urge researchers to continue to investigate anger reactivity as influencing the development of behavior problems across multiple levels of analysis (e.g., molecular, neural, emotional, behavioral, environmental) transactionally in an effort to address this important area of developmental psychopathology research.

Surgency Whereas inhibited children are biologically predisposed to experience high negative affect in response to novelty (García Coll et al., 1984; Kagan, 1997), another temperamental superfactor, surgency, and associated temperament type, exuberant, that is characterized by approach, rather than withdrawal, to novelty, and the high levels of positive affect, activity, and impulsivity that accompany approach has been identified. Historically, temperament research focused largely on negative emotionality, including fear and anger reactivity, difficult temperament, and behavioral inhibition. More recently, temperament constructs including positive reactivity and approach have become of interest to more researchers. This growing interest has been attributed at least in part to an understanding that some children are not simply low in behavioral inhibition but instead take pleasure encountering novelty (PolakToste & Gunnar, 2006). As noted earlier, varying research approaches have examined and defined the characteristics of these children. Rothbart's surgency superfactor has been used to label variations in approach to novelty (Gartstein & Rothbart, 2003; Rothbart, Ahadi, et al., 2001). Categories of children with these characteristics may also be labeled uninhibited or exuberant, depending on the methodology and/or perspective employed (Calkins et al., 1996; García Coll et al., 1984; Putnam & Stifter, 2005; Rothbart, Ahadi, et al., 2001). For the

purposes of this literature review, we use the terms that were used in the studies reviewed but consider them very similar in conceptualization. The growing area of research on surgent, uninhibited, and exuberant children has indicated that although these children display high levels of positive affect, which is typically thought to be an adaptive quality, they are at risk for developing later behavior problems, within the realms of both externalizing and internalizing behaviors (Putnam & Stifter, 2005; Rubin, Coplan, Fox, & Calkins, 1995; Schwartz, Snidman, & Kagan, 1996; Stifter et al., 2008). Concurrent associations between exuberance/surgency and externalizing behavior problems have been found in toddlers (Gartstein et al., 2012; Putnam & Stifter, 2005), elementary school children (Oldehinkel et al., 2004; Rothbart, Ahadi, et al., 1994), and adolescents (Muris, van der Pennen, Sigmond, & Mayer, 2008). There is also evidence for a longitudinal association between exuberance and externalizing behavior problems (Rothbart, Ahadi, et al., 2000). Stifter, Putnam, & Jahromi (2008) found that children identified as exuberant at age 2 were more likely to be rated as higher in externalizing behavior problems at age 4 than other children. Dollar and Stifter (2012) found that highly surgent children were more likely to show negative peer behaviors observed in the laboratory as well as higher parent-rated aggression than children lower in surgency longitudinally. In addition, Degnan and colleagues (2011) reported that a high stable exuberance profile over toddlerhood and early childhood was associated with higher ratings of 5-year externalizing behavior and disruptive behavior. This association was especially strong for exuberant children who showed left frontal electroencephalographic (EEG) asymmetry. Exuberant children may also be at risk for developing internalizing behavior problems; however, this evidence is mixed. In one study (Stifter et al., 2008), exuberant children were not only more likely to be rated as high in externalizing behavior problems but also higher in internalizing behaviors than inhibited children. Some studies, however, suggest that exuberance may offset the risk for internalizing behavior problems. For example, low exuberance has been linked to depressive and anxiety disorders in toddlerhood (Dougherty et al., 2011). There are theoretical explanations for each of these associations, and the mixed findings could be the result of differences in measurement. As this area of research is relatively new, additional research examining the associations between exuberance the development of internalizing and externalizing behavior problems is warranted. In addition, research examining specific types of externalizing behaviors that exuberant children are at risk for developing is greatly needed. There has been a recent surge of interest in the developmental trajectories of surgent/exuberant children; yet there still is a dearth of research examining the moderating role of environmental factors, such as parenting behaviors, in affecting children's later psychopathology. The limited research that exists has largely examined the development of behaviors and deficits in skills associated with psychopathology. Root and Stifter (2010) found that high-approach children (similar to children high in surgency) showed higher levels of teacher-rated disruptive behavior in the classroom when they had mothers who reported high levels of punitive, dismissive responses to their children's negative emotions. Also, as previously mentioned, Kochanska and colleagues (Kochanska, 1991, 1995, 1997; Kochanska, Aksan, & Joy, 2007) have extensively examined the interaction between parenting behaviors and fearless children's

(similar to exuberant children) behaviors in predicting the development moral and behavioral self-regulation, factors related to problem behavior (Calkins & Keane, 2004; Keenan & Shaw, 2003). Kochanska (1991, 1995) hypothesized that because the fear threshold is higher in fearless children, as compared to fearful children, gentle discipline would not induce enough fear arousal to foster self-regulated compliant behavior. However, simply increasing the applied pressure with these children, termed power assertiveness, was not associated with internalization of rules and standards for fearless children. Rather, Kochanska (1997) found that a mutually positive orientation reflected in shared positive affect or a secure attachment that optimizes positive motivation was more effective at promoting internalization and compliance with fearless children. Additional support has been found for the important role of parental warmth in promoting exuberant children's self-regulatory abilities in research showing that exuberant children whose mothers used commands and prohibitive statements coupled with a positive tone were more likely to show effortful control later in childhood (Cipriano & Stifter, 2010). Similarly, the development of self-regulation in exuberant children may depend on maternal attention-directing behaviors. Conway and Stifter (2012) found that exuberant children of mothers who helped themmaintain their attention during a problem-solving task developed better inhibitory control than exuberant children whose mothers did not use this attention strategy. Thus, early parenting behaviors appear to moderate child exuberance to predict important regulatory behaviors that protect against the development of problem behavior. Although this research is an excellent start in identifying important developmental processes in exuberant/surgent children, more research is needed to examine how these processes work together to affect development of behavior problems in these children.

Positive Emotionality Despite the focus on negative emotions historically, interest in temperamental positive emotionality has grown and appears to be a promising link toward a greater understanding of children's trajectories toward mental health or alternatively dysfunction. In general, a predisposition toward childhood positive emotions, generally measured in terms of happy, cheerful mood, is thought to be adaptive, as it has been found to be related to peer competence and prosocial behavior (Denham, McKinley, Couchoud, & Holt, 1990; Eisenberg et al., 1996). Indeed, there is considerable research linking positive emotionality to adaptive developmental outcomes and lowered risk for psychopathology. For example, one study found that lowintensity positive emotionality exhibited during quiet and interpersonal situations was related to the ability to self-regulate emotion and behavior (Kochanska, Aksan, Penney, et al., 2007) while another study showed parent ratings of low-intensity pleasure to be negatively associated with both internalizing and externalizing behavior problems (Gartstein et al., 2012). Additional support for the importance of positive emotionality in lowering risk of psychopathology comes from an expanding literature showing that low positive emotionality, as early as toddlerhood, is a risk factor in the development of depression. For example, low levels of positive emotionality in children have been linked with EEG asymmetries believed to characterize depressed adolescents and adults as well as to a history of depressive disorders in their mothers (Durbin, Klein, Hayden, Buckley, & Moerk, 2005; Shankman et al., 2005).

Further, Dougherty and colleagues (2010) found that low positive emotionality at age 3 predicted depressive symptoms at age 10, after controlling for negative emotionality and depressive behaviors at age 3. Interestingly, these authors also reported that negative emotionality predicted depression only when children were also low in positive emotionality. Although there is theoretical and empirical work outlining how positive emotionality in children promotes them as attractive social partners and serves as a protective factor for children in the development of depression, positive emotionality also has been linked to maladaptive social and psychological outcomes (Putnam, 2012). For example, high-intensity pleasure has been linked to expression of anger/frustration from infancy through childhood (Putnam, Rothbart, & Gartstein, 2008; Rothbart, Ahadi, et al., 1994; Rothbart, Derryberry, et al., 2000). Further, empirical work has shown that children who show more intense joy and are quicker to show joy score lower on effortful control measures (Kochanska et al., 2000); laboratory observations of smiling and laughter predict later impulsivity and lower inhibitory control (Rothbart, Derryberry, & Posner, 1994). More specific to problem behaviors, Rothbart and colleagues (1994) found that infant smiling and laughter was related to parent-reported aggression at age 7. Rydell and colleagues (2003) reported that 5- to 8-year old children's positive emotionality/exuberance was related to maternal reports of children's externalizing behavior problems, whereas children's ability to regulate positive emotionality/exuberance was associated with lower externalizing behavior problems. It is important to note that positive emotionality, particularly at high levels of intensity, is a central defining characteristic of temperamental exuberance and surgency which has been linked repeatedly to the development of aggressive, externalizing behavior problems (Degnan et al., 2011; Stifter et al., 2008). It may be that the other characteristics of exuberance, such as approach to novelty, impulsivity, and activity level, together with high-intensity positive affect, contribute toward this risk for problem behavior rather than positive affect alone. Although the area of inquiry regarding the developmental role of positive emotionality in children's later behavior problems is still in its early stages, there is some support for the role of environmental factors as moderating the association between positive emotionality and later behavior problems. For example, there is some evidence that low positive emotionality may be related to risk of depression and other behavior problems through environmental factors such as lower parental support, acceptance, and involvement (Branje, van Lieshout, & van Aken, 2005; Lengua & Kovacs, 2005). Further, maternal rejection has been shown to be more strongly associated with children's depression and conduct problems for children who were low in positive emotionality as compared to those who were moderate and high in positive emotionality (Lengua, Wolchik, Sandler, & West, 2000). Considered conversely, these results suggest that positive emotionality may be a protective factor for children in negative environmental contexts by buffering the harmful effects of less-than-optimal parenting (Lengua & Wachs, 2012).

Activity Level Activity level has consistently been considered an important temperament trait (A. Buss & Plomin, 1975; Henderson & Wachs, 2007; Thomas & Chess, 1977). There are two central

components to activity level, tempo and vigor, and it is commonly measured as the duration of energetic behavior, displacement of body movements, and rate and amplitude of speech and movement (A. Buss & Plomin, 1975). Generally, high activity is associated with externalizing behavior problems generally, as well as ADHD difficulties (De Pauw et al., 2009). Indeed, the relationship between temperamental activity level and ADHD is well established (J. White, 1999). For example, children with ADHD are rated significantly higher on activity than community children (De Pauw & Mervielde, 2010; Foley, McClowry, & Castellanos, 2008). Foley, McClowry, and Castellanos (2008) examined the associations between ADHD and child temperament dimensions in two groups of 6- to 11-year-old children: one group of children showing symptoms of ADHD and another group of children exhibiting no symptoms of ADHD. Children exhibiting high levels of ADHD symptoms (hyperactivity, impulsivity, and inattention) were rated as significantly higher in activity level than those with no ADHD symptoms. More specifically, temperamental activity was found to be related to the hyperactive-impulsive and combined types of ADHD but not the inattentive type (Stringaris, Maughan, & Goodman, 2010). Activity level has also been considered as a core component of the surgency superfactor (e.g., Rothbart, Ahadi, et al., 2001). Indeed, activity level may underlie the link between surgency and externalizing behaviors. And, while high levels of activity characterize children with hyperactivity problems, it also characterizes children with disruptive behaviors (Nigg, 2006). However, children with high levels of activity are also likely to be more impulsive and distractible, which might blur the relation between activity and the development of ADHD and/or externalizing behavior problems. Summary Historically, temperament research focused largely on negative emotionality, including fear and anger reactivity, difficult temperament, and behavioral inhibition. More recently, temperament constructs, including positive reactivity and approach, have been of interest to more researchers. Although still a relatively new phenomenon, it is now acknowledged that exuberance is not simply the “opposite” of behavioral inhibition, and research is needed to identify the unique behavioral and neurobiological qualities of these children. Children who are more likely to approach novelty, frequently labeled uninhibited, exuberant, or surgent, are characterized as high in positive affect, activity level, and impulsivity and low in shyness and withdrawal. Although these children display high levels of positive affect, which is typically thought to be an adaptive quality, they are at risk for developing later behavior problems. Positive emotionality has also been considered as a separate temperament trait, not only as a component of temperamental surgency or exuberance. Indeed, multiple lines of research have shown the adaptive role of positive emotionality, including the development of self-regulation and lowering risk of depression. Yet the role of positive emotionality in children's developmental trajectories is not entirely clear as there is additional research linking this construct with maladaptive behaviors, such as externalizing behaviors and low self-regulation Low positive emotionality appears to be a risk factor for depression and other behavior problems, particularly when combined with negative parenting behaviors, such as low

acceptance, and low involvement and negative parenting behavior may exacerbate the risk for psychopathology for children low in positive emotionality. Investigations on the role of surgency, exuberance, positive emotionality, and activity level in relation to emotional and behavioral disorders are just now emerging, and the results are promising. Additional research examining the associations between exuberance/surgency and the development of specific types of externalizing behaviors, however, is greatly needed, as well as identification of which children are at risk for developing comorbid internalizing and externalizing behavior problems. Although there has been a surge of interest in the developmental trajectories of exuberant children, there is still a lack of understanding regarding the mechanisms involved in risk or protective factors for these children, including both environmental and innate factors of the child. Further, there has been a dearth of research examining the interaction between children's exuberance and their environment in the development of behavior problems. Finally, although there is some support for factors associated with lowering risk of developing behavior problems for exuberant children such as effortful control ), research in temperamental exuberance is still very much in its infancy, and many questions still remain. In addition, there is still much work needed regarding the role of positive emotionality as a protective or risk factor in children's development of behavior problems. Recent reviews of the research literature on temperamental positive emotionality have highlighted the importance of considering the type and/or intensity of positive emotion in understanding its relation in children's adjustment. Specifically, positive emotionality that is associated with strong approach behavior and/or is experienced in high-intensity contexts is hypothesized to be associated with problematic behaviors in the externalizing realm, whereas low-intensity positive emotions, such as contentment and affiliation, are more likely to be associated with the development of self-regulation and desirable social behaviors. More research to support this position, however, is needed before the risk and protective qualities of positive emotionality are fully understood.

Effortful Control The construct of effortful control was introduced by Rothbart and colleagues (Rothbart, 1989; Rothbart & Ahadi, 1994; Rothbart & Bates, 2006) to describe the self-regulatory components of temperament. Effortful control is defined as the ability to inhibit a dominant response in order to perform a subdominant response, to detect errors, and to engage in planning (Rothbart & Bates, 2006; Rudea, 2012) and includes the dimensions of inhibitory control and attentional focusing/control. Because effortful control is voluntary and active, this temperament construct is differentiated from more reactive forms of regulation, fear reactivity and impulsivity (Derryberry & Rothbart, 1997). Effortful control is hypothesized to affect children's developmental trajectories by contributing to their ability to process information and modulate emotion and behavior. Whereas negative emotionality and surgency are associated with behavior problems, effortful control as a temperament factor is largely considered to protect children against maladaptive outcomes, including the development of psychopathology. Indeed, a high level of effortful control is the one temperament trait that has been most often linked to

behavior problems by lowering risk (e.g., Eisenberg, Spinrad, & Eggum, 2010), and the evidence is quite strong and consistent (Eisenberg, Sadovsky, et al., 2005; Lengua, 2006; Lengua et al., 1998; Oldehinkel, Hartman, Ferdinand, Verhulst, & Ormel, 2007; Olson, Sameroff, Kerr, Lopez, & Wellman, 2005; Spinrad et al., 2007; Valiente et al., 2003). There is also evidence that the role of effortful control as a protective factor persists into childhood (Muris et al., 2008) and adolescence (L. Ellis, Rothbart, & Posner, 2004) by being negatively associated with anxiety, depression, and aggression. Although low levels of effortful control are related to a host of problematic behaviors, empirical evidence suggests that it is most strongly related to the development of externalizing behavior problems, even after controlling for other cognitive and social risk factors (Olson et al., 2005; Valiente et al., 2003). For example, children with externalizing behavior problems were reported to have deficits in attentional and effortful control, in addition to high impulsivity, as compared to children with internalizing behaviors and control children (Eisenberg et al., 2009). Deficiencies in effortful control have also been observed in children with ADHD (see Barkley, 1997; Nigg, 2001; van der Meere 2002), a clinical disorder often associated with externalizing behavior problems. In a large-scale study of preadolescent temperament and problem behaviors, Oldehinkel et al. (2004) reported that low effortful control was the central temperament predictor of externalizing behavior problems, reporting an effect size greater than .50. Importantly, although this study used parent ratings for both temperament and behavior problems, confounding items were removed, and the strength of the relationship between effortful control and externalizing problems was not diminished, nor did it change when the researchers examined adolescents self-ratings of externalizing behavior. The relationship between low effortful control and externalizing behavior problems has been found across time with the stability of relations also taken into account, indicating that change in effortful control is related to change in children's externalizing behavior problems (Belsky, Pasco Fearon, & Bell, 2007; Valiente et al., 2006). For example, attentional dysregulation predicted externalizing behavior problems at the next assessment for children at 54 months, first grade, and fifth grade, while controlling for previous levels of these constructs (Belsky, Pasco Fearon, et al., 2007). However, at ages 18, 30, and 42 months, researchers found no evidence for a causal relation between effortful control at one time point and externalizing behavior problems at a later time point. Instead, externalizing behavior problems preceded lower levels of effortful control, suggesting that externalizing behaviors may counteract any growth in regulatory abilities (Eisenberg, Spinrad, Eggum, et al., 2010). Effortful control has also been linked to internalizing behavior problems; however, the association between these two constructs is not always clear. Some researchers have found that effortful control is negatively related to internalizing behavior problems (Eisenberg et al., 2001; Lengua, 2006; Martel et al., 2007; Oldehinkel et al., 2007). For example, parent report of low effortful control and observed attentional control during middle childhood was found to be related to greater internalizing behavior problems (Lemery-Chalfant, Doelger, & Goldsmith, 2008) whereas increases in effortful control over a 2-year period was associated with fewer internalizing behavior problems in preadolescence (Lengua, 2006). In contrast, others have found a positive relation between effortful control and internalizing behavior problems

(Murray & Kochanska, 2002; Rydell et al., 2003), suggesting issues of overcontrol. Adding to the complexity of the relation between effortful control and internalizing problems are findings of a quadratic relation between effortful control and internalizing behavior problems (Murray & Kochanska, 2002). Specifically, in a normative sample, a quadratic relation was revealed between preschoolers' observed effortful control and parent-rated total problem behaviors (combined internalizing and externalizing behaviors). Follow-up analyses suggested that this association was because children with higher effortful control were rated by their mothers as higher in internalizing behavior problems than children with moderate levels of effortful control. Some researchers have hypothesized that the more specific behaviors that make up effortful control (i.e., attention shifting/control, inhibitory control) might account for the mixed findings (Eisenberg et al., 2009; Klein, Dyson, Kujawa, & Kotov, 2012). As with the other temperament superfactors (e.g., negative emotionality, surgency), effortful control is heterogeneous, and the association between it and internalizing behaviors may depend on how effortful control is measured as well as the other innate aspects of the child. A study by Gartstein and colleagues (2012) found that attentional control was primarily responsible for the negative association between effortful control and internalizing behavior problems, especially given the lack of significant associations between inhibitory control and internalizing behavior problems. Interestingly, low attentional and inhibitory control have been associated with externalizing behavior problems in the toddler and preschool years (Gartstein et al., 2012; Hill, Degnan, Calkins, & Keane, 2006; Lemery et al., 2002; Spinrad et al., 2007) as well as later in childhood and preadolescence (Belsky, Bakermans-Kranenburg, et al., 2007; Eisenberg, Sadovsky, et al., 2005; Eisenberg et al., 2009). Thus, it appears that it is necessary to take a more fine-grained approach to examining the link between effortful control and developmental psychopathology. The development of effortful control is very likely to be susceptible to environmental influences and parenting, especially during the preschool years, when children's development of these abilities is growing rapidly (Lengua, Honorado, & Bush, 2007; Rothbart & Rueda, 2005). Further, effortful control may serve as a protective factor against maladaptive environmental influences, where children low in effortful control who also experience lessthan-optimal environments may be at a greater risk of developing behavior problems than children with high levels of effortful control. Morris and colleagues (2002) found that there was an especially strong relation between mother hostility and children's externalizing behavior problems for children rated as low in effortful control. For children low in effortful control, positive parental control appears to buffer this risk (Karreman, van Tuijl, van Aken, & Deković, 2009). Eisenberg, Cumberland, and Spinrad (1998) hypothesized that parents who are sensitive, warm, and express appropriate emotions are more likely to raise children with higher levels of effortful control, thereby lowering their chances of developing behavior problems. Empirical evidence has been found for this proposal by showing that children's effortful control mediates the relation between parenting behaviors and children's development of behavior problems. For example, the relation between maternal expressivity and later internalizing and

externalizing behavior problems was mediated by children's effortful control (Valiente et al., 2006). In addition, Spinrad and colleagues (2007) found that concurrently at both 18 and 30 months, children's effortful control mediated the association between supportive parenting and externalizing problems. Evidence supporting the mediating role of effortful control in the association between various environmental factors and later behavior problems has also emerged. For example, chaos in the home when children were 3 years old was indirectly related to externalizing behavior problems 2.5 years later through children's inhibitory control at age 4 (Hardaway, Wilson, Shaw, & Dishion, 2012). Summary Effortful control, the self-regulatory aspect of temperament, defined as the ability to inhibit a dominant response in order to perform a subdominant response, to detect errors, and to engage in planning, is comprised of inhibitory control and attentional control. Effortful control is hypothesized to affect children's developmental trajectories by contributing to their ability to process information and modulate emotion and behavior. Whereas many other temperament traits are positively associated with behavior problems and psychopathology, effortful control is largely considered a protective factor against maladaptive outcomes. Indeed, there is extensive evidence to suggest that effortful control is negatively related to children's development of behavior problems, and this association is especially strong for externalizing problems. Effortful control has also been linked to internalizing behavior problems, but this association is not always clear. The research also indicates that effortful control may serve as a protective factor against maladaptive environmental influences, where children high in effortful control who experience less-than-optimal environmental experiences may be buffered against developing behavior problems and psychopathology. Although the research indicates that effortful control has a central role in maintaining a positive developmental trajectory, how this ability develops in light of the child's level and type of temperamental reactivity and what supports are needed for its development are still unclear. Several studies have indicated that parents are especially important to the child's developing self-regulation when these parenting behaviors vary according to the child's temperament. Research linking these processes to indices of behavioral adjustment and maladjustment are needed.

Temperament × Temperament Interactions There are numerous possible patterns by which one temperament trait moderates the association between another temperament trait and children's psychopathology; however, although the amount of research is growing, relatively little has been conducted on this area of inquiry. Here we provide an overview of existing research on temperament × temperament interactions in predicting children's behavior problems.

Effortful Control × Temperament Although research examining temperament × temperament interactions is still limited, there are a number of studies examining the moderating role of effortful control on the association

between other temperament traits and developmental psychopathology (Rothbart & Bates, 2006). Indeed, effortful control commonly interacts with other temperament dimensions, such as negative emotionality and impulsivity, to predict behavior problems, especially within the externalizing realm (Eisenberg et al., 1996, 2004; Valiente et al., 2003). This association is to be expected since it is widely hypothesized that the purpose of the regulatory aspects of temperament is to modulate the reactive aspects of temperament—especially because, as outlined earlier, unregulated reactivity frequently leads to maladjustment (Rothbart & Bates, 2006). This position is supported by a number of studies. For example, Muris, Meesters, and Blijlevens (2007) found that effortful control weakened the positive relation between negative emotionality and both internalizing and externalizing behavior problems. Likewise, Lonigan, Vasey, Phillips, and Hazen (2004) reported that highly negative children were at risk for developing anxiety difficulties only when they were also low in self-regulation. Researchers have also found that a negative association between effortful control and externalizing behavior problems was strongest for children with high levels of negative emotionality (Valiente et al., 2003); temperamentally negative children who exhibited better effortful control were less likely to develop behavior problems 4 years later. The moderating role of effortful control on the relation between negative emotionality and externalizing behavior problems has also been found with pre- and early adolescents (Oldehinkel et al., 2007). Studies examining more specific forms of temperamental reactivity found that infants and toddlers high in anger and activity level and low in effortful control (measured with executive function tasks) were more likely to develop ADHD (Auerbach et al., 2008; Willcutt, Doyle, Nigg, Faraone, & Pennington, 2005). Additional support for interactive effect of anger and effortful control has come from a short-term longitudinal study of preschoolers (Diener & Kim, 2004). Results from this study indicated that children whose mothers rated them as easily frustrated and low on self-regulation (composite of low inhibitory control and high impulsivity) were rated by their teachers as being high in externalizing behaviors. Components of effortful control have also been investigated for their role in mediating or moderating the association between other temperament traits and the development of behavior problems. Sustained attentive behavior was found to moderate the association between temperamental anger and aggressive behavior problems in children (Kim & Deater-Deckard, 2011). Evidence for this association has also been found in adolescence (Oldehinkel et al., 2004), 2004). Deater-Deckard, Petrill, and Thompson (2007) found that the link between temperamental anger and later aggressive behavior was mediated by children's regulation of sustained attentive behavior, indicating a lack of effortful regulation. As can be seen, there is good evidence supporting the importance of effortful control in altering the association between anger/negative emotionality and externalizing behaviors; however, it is important to note that other researchers have failed to find evidence of these effects (Belsky, Friedman, & Hsieh, 2001; Rydell et al., 2003). The moderating role of effortful control is also believed to be important for children at risk for developing internalizing behavior problems, although the association is not as strong as with externalizing behavior problems. Even though there are significant associations between these variables, the role of effortful control in moderating the association between fear/behavioral

inhibition and internalizing behavior problems is mixed. Some research has indicated the protective role of high effortful control in lowering risk of developing internalizing behaviors. Oldehinkel and colleagues (2004) showed that adolescents displaying high fear and shyness and moderately low effortful control were high in internalizing behavior problems. Additional support for the moderating role of effortful control has come from Eisenberg, Shepard, Fabes, Murphy, and Gutherie (1998), who found that shy children who were low in attention shifting had greater internalizing symptoms, whereas children who were better at attention shifting showed discontinuity in these behaviors over time. However, other studies suggest that inhibitory control, defined as the ability to effortfully inhibit behavior when necessary (Rothbart & Bates, 2006), is not always adaptive, particularly for behaviorally inhibited children. Children high in behavioral inhibition who also showed greater inhibitory control were more likely to be rated higher in social anxiety by their teachers (Thorell, Bohlin, & Rydell, 2004) and were more socially withdrawn (Fox & Henderson, 1999) than behaviorally inhibited children low in inhibitory control. More recently though, greater use of attention shifting was associated with a reduced risk of anxiety for behaviorally inhibited children whereas high levels of inhibitory control were associated with increased risk of anxiety (L. White, McDermott, Degnan, Henderson, & Fox, 2011). The authors and others have proposed that it may be critical for behaviorally inhibited children to be able to flexibly shift their attention to help regulate the negative emotions they experience in certain situations. Whereas inhibitory control may serve as an important regulatory ability for most children, heightened inhibitory control may be detrimental for behaviorally inhibited children. Behaving in a rigid, inflexible, and overly cautious manner in many situations may heighten their risk for anxious behaviors and hinder their chances of developing socially appropriate behaviors. In sum, effortful control may serve a different role for behaviorally inhibited children in increasing or lowering risk of developing internalizing behavior problems, depending on how and which subcomponent of effortful control (attentional control, inhibitory control) is measured in the study. This is an interesting and important avenue for future research. Research identifying the mechanisms underlying the relation among attention shifting, inhibitory control, and internalizing behavior problems, particularly in behaviorally inhibited children, is needed. The work of Perez-Edgar (Perez-Edgar et al., 2010; Perez-Edgar et al., 2010; Perez-Edgar et al., 2007) on the neural and genetic underpinnings of attention bias in behaviorally inhibited children who develop anxiety highlights a promising area for further study. Likewise, examining the developmental trajectories of both attentional control and inhibitory control in behaviorally inhibited children may increase our understanding of when and how these behaviors develop into anxiety.

Surgency × Temperament Until recently, we knew very little about the explanatory mechanisms involved in exuberant children's development of aggressive and externalizing behaviors. Although few studies exist in this area, one leading hypothesis points to the role of exuberant children's predisposition toward anger reactivity. Because limits are frequently placed on attempts to approach aspects

of their environment, exuberant children are more likely to experience intense reactions of anger when their goals are blocked (Derryberry & Reed, 1994; Rothbart & Bates, 2006; Rothbart, Derryberry, et al., 2000). Indeed, past research has shown that the greater the interest or positive affect experienced when engaging with the environment (e.g., toy, mobile, prize), the more anger expressed when the stimulus was not obtainable (Dennis, 2006; Lewis, WolanSullivan, & Ramsay, 1992). Consequently, early confrontations in which rules of behavior are required may provide important opportunities for exuberant children to learn emotion regulation, a form of self-regulation. If they have not acquired this ability by childhood, exuberant children may be at risk for negative outcomes, such as externalizing behavior problems, as their inability to regulate emotions may cause them to act inappropriately in certain situations, even though at other times they are highly sociable and outgoing (PolakToste & Gunnar, 2006). In support of this position, Stifter and colleagues (2008) found that exuberant children who displayed higher levels of negative emotion and lower levels of positive/neutral emotion to a disappointing situation were rated by their parents as having higher levels of externalizing and total problem behaviors than exuberant children who could regulate their emotional expression. Additional research has found that although highly surgent children are at risk for developing aggressive behaviors, the ability to use specific emotion regulation behaviors in a frustrating situation lowered this risk (Dollar & Stifter, 2012). In another study, children rated as high in surgency and low in self-regulation have been found to display higher levels of aggressive behaviors, which in turn was associated with peer rejection (Gunnar, Sebanc, Tout, Donzella, & van Dulmen, 2003). Taken together, these findings indicate that the development of emotion regulation, especially the ability to regulate anger, is a critical developmental task for exuberant children. Neurobiological models of approach (e.g., Depue & Collins, 1999; Eysenck, 1967; Gray, 1987) suggest that high levels of approach behavior, such as that exhibited by exuberant children, must be offset by appropriate inhibitory mechanisms. Based on these models, exuberant children are hypothesized to require the ability to regulate their behavior as well as their emotions. Recent evidence suggests that aspects of developing effortful control processes are critical for exuberant children. Lahat and colleagues (2013) found that exuberant children were significantly more likely to show an increased risk-taking propensity, which is associated with various aspects of psychopathology, when they were low in attention-shifting abilities. Finally, we have found initial evidence that although positive affect, even at intense levels, can be an adaptive trait for exuberant children, the combination of intense positive affect and high activity level in an engaging task put these children at risk for developing conduct problems (Dollar, 2011). These findings suggest that for exuberant children, intense positive affect when accompanied by other “disruptive” behaviors may be a marker for developing externalizing behaviors, likely because their intense excitement becomes too active and impulsive to control. Thus, in addition to learning to regulate anger, it is adaptive for exuberant children to develop the ability to regulate their often impulsive, active behavior.

Positive Emotionality × Temperament

The growth in research examining the role of positive emotionality in the development of psychopathology has produced several studies focused on the interaction between positive emotionality and other temperament traits. One study has shown that for toddlers low in negative emotionality, low positive emotionality was associated with higher rates of parental depression, a known risk factor for childhood depression (Olino, Klein, Dyson, Rose, & Durbin, 2010). Additional support for the interactive role of positive emotionality and negative emotionality in predicting later depression has been found in research on asymmetry in resting EEG activity in frontal and posterior regions, which are putative biomarkers for depression. Specifically, results from this study found that children with low positive emotionality or high negative emotionality were at risk for depression (EEG asymmetry was used as a marker for depression) and that children low in positive emotionality and high in negative emotionality were not at a greater risk than those children with only one of these traits (Shankman et al., 2011). It is noteworthy that although low positive emotionality has been linked to the development of depression, more recent research has found a temperament × temperament interaction for anxiety. Behaviorally inhibited children who exhibited low positive emotionality were more likely to be diagnosed with anxiety (Dougherty et al., 2013). More recently, we have been interested in the moderating role of activity level on the relation between individual differences in positive emotionality and the development of behavior problems (Dollar, 2011). Children who showed low activity and low positive affect when playing a game with their mothers were rated as high in social withdrawal; higher levels of positive affect for children low in activity significantly reduced the likelihood of being socially withdrawn. Although corresponding with much extant research showing that children low in approach and activity but high in negative affect and withdrawal are at risk for being socially withdrawn (García Coll et al., 1984), our findings considered positive as well as negative affect. Because much research has focused on the importance of children at risk for developing internalizing behaviors and social withdrawal to learn to regulate their fear (K. Buss, 2011; Fox, 1994), it is important that additional research substantiate these findings that increasing positive affect may protect children, particularly behaviorally inhibited children, from developing these problem behaviors. Summary An important area of temperament research that is slowly accumulating empirical evidence is that examining the association between temperament and psychopathology within the context of other temperament traits, or temperament × temperament interactions. Although there are numerous possible patterns by which one temperament trait moderates the association between another temperament trait and children's psychopathology, few studies have examined this area of inquiry. One exception is a fairly extensive literature outlining the importance of effortful control in lowering highly negative children's risk of developing behavior problems; this association is especially true within the externalizing realm. Further, there is much support for the interactive and additive effects of anger with other temperament traits, especially low levels of self-regulatory abilities, in predicting behavior problems. Additional research has considered how effortful control lowers the risk of developing anxiety problems for

behavioral inhibited children. However, the findings indicate that different components of effortful control, attentional shifting and inhibitory control, are differentially associated with increasing or lowering the risk of developing anxious behaviors. Another growing area of temperament × temperament research considers the developmental trajectories of children varying on their levels of surgency. Although this research is still in its infancy, one leading hypothesis points to the importance of anger regulation for highly surgent children in lowering their risk of developing aggressive, externalizing behaviors. Other recent research has presented support for attentional regulation as a critical developmental ability for surgent children to offset their risky behaviors. Another critical ability for exuberant children that has received recent empirical support is the capacity to regulate highly active, impulsive behavior when excited while preserving the positive affect, even at intense levels. Finally, although there has been limited research examining the role of positive emotionality, recent research points to the importance of considering it within the context of the related temperamental constructs of negative emotionality, behavioral inhibition, and activity level.

Neurobiology Linking Temperament and Developmental Psychopathology Contemporary theories of temperament emphasize a biological or genetic basis for temperament and often include it as part of the definition (Rothbart & Derryberry, 1981). The central tenet of A. Buss and Plomin's (1975, 1984) theory, for example, is that temperament is largely heritable, and behaviors must be present early in life to be considered a temperament trait. The work of Goldsmith and Campos (e.g., Goldsmith & Campos, 1982; Goldsmith, Campos & Lemery, 1997; Lemery et al., 2002) also considers early-appearing individual differences primarily in emotions and emotion regulation and provides empirical support from behavioral genetics research on temperament. Likewise, the research of Kagan and colleagues (e.g., Kagan, 2002; Kagan, Reznick, Clarke, & Snidman, 1984; Kagan et al., 1988) on behavioral inhibition is strongly rooted in the premise that a child's heritable neurochemistry affects arousal in the amygdala and/or the bed nucleus of the stria terminalis and their projections, which is reflected in the child's withdrawal from novelty. Finally, Rothbart (Rothbart, 1981; Rothbart & Ahadi, 1994; Rothbart & Derryberry, 1981) has long argued that an integrated view of temperament requires temperament research to be conducted at the genetic, neural, endocrine, autonomic, and central nervous systems levels. Rothbart's model emphasizes the autonomic basis of both reactivity and regulation, in addition to the maturation of neutral pathways and brain structures associated with the development of children's regulatory behaviors. The theoretical links between temperament and psychopathology, particularly the spectrum model, are strengthened by studies examining their shared neurobiological bases (Goldsmith & Lemery, 2000; Nigg, 2006). Alternatively, variations in biological functioning associated with temperament may make children more vulnerable to negative or aversive environments resulting in behavioral and emotional disorders which would support the

vulnerability/predisposition model linking temperament to psychopathology (McCrory, DeBrito, & Viding, 2010). Studies testing the differential susceptibility hypothesis, for example, have shown that candidate genes associated with temperament, such as DRD4), interacts with various forms of parenting to predict maladaptive behavior (BakermansKranenburg & van IJzensoorn, 2006). Unfortunately, to date, studies such as these are rare. Although theories of temperament are based on the notion that individual differences in temperament are biologically based, it is critical that the environment is also acknowledged as serving an important role in the expression of temperament (Goldsmith et al., 1987). Recent perspectives have acknowledged that there is a transaction between biological processes and the environment beginning prior to birth and occurring throughout development, suggesting that temperament is a complex process of development (e.g., Shiner et al., 2012). For example, temperament is not only heritable but is affected by genetics later in development (Saudino & Wang, 2012). This review highlights the various levels of biological functioning that underlie temperament and its link to developmental psychopathology, but as outlined throughout, we encourage the reader to consider the joint workings of biology and the environment in affecting children's development.

Cardiac Measures Assessments of cardiac functioning have a long history of serving as central biological measures applied to the study of temperament, and the research to date indicates that cardiac measures account for some of the variation in children's observed temperament. The most popular cardiac measures of heart rate, heart rate variability, RSA, and pre-ejection period (PEP), reflect to varying degrees activity in the autonomic nervous system (ANS). The ANS, primarily responsible for the physiological arousal related to emotions, is broken down into two branches: the PNS and the sympathetic nervous system (SNS), which function synergistically. The PNS is predominantly active during moments of relative calm and functions to sustain homeostasis for the individual. The SNS, however, is active during times of stress and produces elevated physiological arousal to assist the individual to respond to physical, psychological, and social challenges. Research on cardiac measures and temperament has focused primarily on measures of the PNS, specifically RSA (sometimes called vagal tone). Porges (2001, 2003; Porges, DoussardRoosevelt, & Maiti, 1994) has put forth the polyvagal theory, which posits that the development of children's ability to regulate their states, motor activity, and emotion is supported by the maturation of the PNS. More specifically, this theory proposes that parasympathetic influence on the heart by way of the vagus nerve enables flexible responses to varying demands of the environment and is therefore considered a marker of physiological regulation (Porges, 2007). Effective vagal functioning is thought to assist in maintaining homeostasis when the individual is faced with situational change or challenges by allowing attention to move from internal processes to environmental demands, thereby supporting social engagement and facilitating the ability to use coping strategies (Porges, 1992, 2007). Studies have demonstrated rather consistently that PNS functioning, as reflected in RSA, is associated with the regulation of attention, emotion, and motor behavior across development (e.g.,

Beauchaine, 2001; Beauchaine, Gatzke-Kopp, & Mead, 2007; Calkins et al., 2002; Porges, 2001, 2007; Porges et al., 1994; Stifter & Corey, 2001; Stifter & Jain, 1996). Two RSA measures are primarily used: baseline RSA and RSA change. Baseline RSA is measured when a person is at rest and denotes the individual's characteristic level of arousal. As such, this measure may reflect temperamental reactivity since it is considered an individual's traitlike ability to engage with the environment. Several studies have demonstrated this relationship (Blandon, Calkins, Keane, & O'Brien, 2010; Calkins, 1997). Higher baseline RSA has been found to be associated with parent and observer reports of greater negative reactivity to anger-eliciting stimuli and positive reactivity to joy-eliciting situations (Calkins, 1997; Calkins et al., 2002; Fox, 1989; Stifter, Fox, & Porges, 1989). Many studies also report high baseline RSA to be associated with other reactive as well as regulatory behaviors, such as approach toward a stranger, activity level, regulated distress in frustrating situations, and lower levels of aggression (Calkins, 1997; Calkins & Dedmon, 2000; Calkins & Fox, 2002; Porges, Doussard-Roosevelt, Portales, & Greenspan, 1996; Stifter & Fox, 1990; Stifter & Jain, 1996). Interestingly, these temperament behaviors are consistent with the surgent/exuberant temperament profile. Also, not surprisingly, several studies have found a link between baseline RSA and behavior problems. Beauchaine and colleagues (2007), for example, found that children diagnosed with clinically high levels of externalizing behavior problems had lower baseline RSA than matched controls. Other researchers have reported low baseline RSA to be related to greater symptoms of anxiety and depression (e.g., Cole, Zahn-Waxler, Fox, Usher, & Welsh, 1996). It is important to note that some studies do not find a relation between baseline RSA and internalizing and externalizing behavior problems (Calkins, Graziano, & Keane, 2007; Fortunato, Gatzke-Kopp, & Ram, 2013; Gentzler, Santucci, Kovacs, & Fox, 2009), suggesting that this relationship may be more complex and likely involves consideration of the context in which these problem behaviors develop. Change in RSA, specifically decreases in RSA in response to a stimulus (vagal withdrawal), is proposed to be a physiological strategy to support regulatory behaviors because it assists the individual to transition from maintaining homeostasis to increasing cardiac output for the purposes of controlling emotional or behavioral responses without engaging the SNS (Calkins, Graziano, et al., 2007; Porges, 2007; Porges et al., 1996; Propper & Moore, 2006). Research examining the relationship between RSA change, temperament, and the development of behavior problems indicates that greater vagal withdrawal is most beneficial for children (Calkins & Dedmon, 2000; Calkins & Keane, 2004; Porges et al., 1996). For example, vagal (RSA) withdrawal during a challenging situation was associated with better regulation (Degangi, Dipietro, Greenspan, & Porges, 1991; Huffman et al., 1998; Stifter & Corey, 2001) and fewer childhood behavior problems (Calkins, 1997; Calkins, Blandon, Williford, & Keane, 2007; Calkins & Keane, 2004; Porges et al., 1996). This may be due in part to the relationship between greater vagal withdrawal and children's effortful control of attention and sustained attention (Calkins, Blandon, et al., 2007; Suess, Porges, & Plude, 1994). However, more recent research has shown that moderate levels of vagal withdrawal and vagal augmentation (increases in RSA from baseline) in some contexts are associated with more adaptive outcomes, such as better executive functioning (Marcovitch et al., 2010), better

academic performance (Graziano & Derefinko, 2013), and social competence (Blair & Peters, 2003). While there has been much research examining the role of PNS functioning in temperament and children's behavioral adjustment, much less work has been conducted regarding the association among various temperament traits, problem behaviors, and cardiac measures of SNS activity. The SNS is associated with the fight-or-flight response to stress (Cannon, 1929). Historically, SNS activity was largely measured by changes in heart rate or heart period (García Coll et al., 1984; Kagan et al., 1984, 1987a), but such measures are influenced by both the PNS and the SNS. For example, increases in heart rate may be due to parasympathetic withdrawal, sympathetic activation, or both. More recently, researchers have focused on PEP. PEP is the sympathetically mediated time between the beat of the heart and ejection of blood into the aorta and has been validated as a measure of sympathetic activity in adults (Berntson et al., 1994) and children and adolescents (Matthews, Salomon, Kenyon, & Allen, 2002; McGrath & O'Brien, 2001; Quigley & Stifter, 2006). Low SNS reactivity, as indexed by longer PEP in response to laboratory challenges, is associated with externalizing behavior problems in childhood (Beauchaine et al., 2007; Boyce et al., 2001; Crowell et al., 2006). Beauchaine, Katkin, Strassberg, and Snarr (2001) also found that decreased PEP reactivity distinguished adolescents comorbid for ADHD and CD from those with ADHD only and from controls. Low SNS activity is hypothesized to be a marker of externalizing disorders because children who are unable to mount a physiological response necessary to experience fear may have deficits in the development of conscience while generating autonomic arousal may protect at-risk children from developing behavioral disorders (Fowles, Kochanska, & Murray, 2000; Frick, Ray, Thornton, & Kahn, 2014; Raine, 1993). Supporting this contention are findings from a study we conducted examining the association between PEP and temperament. Highly surgent children who exhibited both higher sympathetic tone and greater sympathetic reactivity scored lower in problem behavior than those who were less sympathetically reactive (Stifter, Dollar, & Cipriano, 2011). Likewise, PEP measures of increased sympathetic reactivity have been associated with toddler freezing behavior during interactions with a stranger (K. Buss, Davidson, Kalin, & Goldsmith, 2004). As the PNS and the SNS operate together in supporting physiological arousal and responses to stress, models outlining the dynamic between these two systems have been proposed (Berntson, Cacioppo, & Quigley, 1991) and applied to psychopathology (Beauchaine et al., 2001; Berntson & Cacioppo, 2004). To date, only a few studies have examined activity in both the PNS and the ANS in children. For example, children with externalizing problems were found to have low reactivity in both branches (Boyce et al., 2001). Another study that used skin conductance as a measure of SNS activity (El-Sheikh et al., 2009), found that within the context of parental marital conflict children who exhibited co-activation (both PNS and SNS are reactive) or coinhibition (both branches are nonreactive) were more likely to develop externalizing behaviors. One possible mechanism for these findings is the ability to regulate emotion. For example, coinhibition was related to children's inability to regulate their emotion during a disappointment task (Stifter et al., 2011). Although promising, none of the current research has considered temperament and its association with childhood behavior problems.

Following Beauchaine (2001), future research interested in the relationship between temperament and developmental psychopathology would benefit from examining the dynamic between both branches of the ANS.

Electrical Brain Activity Another biological construct relevant to temperament is neural activation. Historically, most of this research has focused on the EEG of infants (e.g., Davidson & Fox, 1982; Fox & Davidson, 1988) and children (Schmidt et al., 1999). Davidson (2000) and Fox (1994) were among the first to investigate how frontal hemispheric asymmetries in EEG were associated with individual differences in temperament and emotional development. Contrary to the notion that left/right-hemispheric activation was related to positive and negative emotions, Davidson and Fox determined that early in development, left activation was associated with approach to novelty while right activation was associated with withdrawal (Davidson, 1992; Davidson & Fox, 1982; Henderson, Fox, & Rubin, 2001). Accordingly, much research has focused on EEG frontal asymmetries in inhibited and exuberant children. Researchers have repeatedly found inhibited children to have greater relative right frontal EEG activity (Fox et al., 1995; Fox, Schmidt, Calkins, Rubin, & Coplan, 1996; L. Schmidt & Fox, 1994). Interestingly, Henderson and colleagues (2001) found that negative reactivity predicted later behavioral wariness only when children also showed right frontal EEG asymmetry in infancy. There is also evidence that left frontal EEG asymmetry lowers inhibited children's risk of continued behavioral inhibition and social wariness (Degnan & Fox, 2007), which in turn may buffer the development of anxiety disorders. Even though most of this research has examined frontal EEG activity for behaviorally inhibited children, there is also some indication that exuberant children show greater relative left frontal activation (Calkins et al., 1996; Fox et al., 2001; Hane, Fox, Henderson, & Marshall, 2008) and that the association between temperamental exuberance and the development of externalizing behaviors may be especially strong for exuberant children who show left frontal asymmetry (Degnan et al., 2011). Event-related potential (ERP), which is computed from ongoing EEG activity, assesses dynamic changes in brain activity by time-locking brain electrical activity when individuals are presented with specific stimuli. As with EEG research, ERP research has examined the association between various ERPs and temperamental fearfulness/inhibition. For example, fearful children of mothers with anxiety disorders showed smaller error-related negativity, a possible biomarker for anxiety, than other children (Torpey et al., 2013). McDermott and colleagues (2009) reported that children who were behaviorally inhibited showed enhanced error-related negativity responses as adolescents, and those behaviorally inhibited children with enhanced error-related negativity responses were at a greater risk of being clinically anxious. It is important to note that other ERP components, such as the N2, also have been shown to be associated with temperamental fearfulness/inhibition (e.g., Perez-Edgar & Fox, 2005b).

Functional Neuroimaging Recent technical advances in neuroimaging have led to a number of studies examining neural

structures and functioning underlying temperamental reactivity and regulation. The majority of these studies were performed with adult subjects, but some have been conducted with adolescents and have provided insight into the specific brain areas associated with temperament. Motivating much of this research is the relationship between temperament and psychopathology and the hypothesis of their shared neurobiological bases. A growing body of empirical work linking temperament and neural functioning has focused on two main areas of the brain: the prefrontal cortex (PFC), including the dorsolateral PFC, the ACC, and the orbitofrontal cortice, and the limbic structures, including the amygdala, hippocampus, and nucleus accumbens (Cloninger, 2000; Kagan, 2002; Posner & Rothbart, 2007; Whittle, Allen, Lubman, & Yucel, 2006). Negative affectivity, specifically fear and anxiety, is associated with activity in the amygdala and the hippocampus while the right dorsolateral PFC and the ACC are involved in the regulation of negative emotions (Whittle et al., 2006). The nucleus accumbens and ACC appear to be involved in positive affectivity. Effortful control/executive function and the attention system are consistently related to activity in the PFC (Posner & Rothbart, 2007). Although most of this research has been done linking adult personality with neural activation, promising new research has been conducted with adolescents and adults identified in childhood as belonging to a particular temperament type. A study of older children and adolescents with aggression (Stadler et al., 2007), for example, showed low activation in the ACC in response to expressions of negative affect compared to controls. Novelty seeking, which characterized the aggressive children, was also related to reduced activity in the ACC. Increased activation of the amygdala to emotion faces in individuals with a history of childhood behavioral inhibition also has received significant attention and empirical support (Perez-Edgar et al., 2007; Schwartz et al., 2003). More specific to developmental psychopathology, a study linking amygdala function, behavioral inhibition, and internalizing disorders (Hardee et al., 2013) revealed amygdala-prefrontal (dorsolateral PFC and anterior insula) connectivity to be negative for angry faces (greater threat) in adolescents previously identified as behaviorally inhibited, which, in turn, was also related to more self-reported internalizing problems. In addition, negative fronto-amygdala connectivity moderated the relation between behavioral inhibition and internalizing disorders. Interestingly, older children and adolescents with callous-unemotional traits (including fearlessness) had reduced amygdala response to fearful faces and lower to negative fronto-amyglada connectivity (Marsh et al., 2008). In summary, research in brain electrical activity and functional neuroimaging has provided some insight into the biology underlying individual differences in reactivity and regulation. However, the bulk of the research has come from studies on behaviorally inhibited children and their risk for anxiety disorders. As exuberant/surgent children are at risk for externalizing problems, more research on the neural correlates of this temperament type, such as the mesolimbic dopamine system, which is involved in reward processing, may give insight to the origins of this risk.

Psychobiological Measures of Stress

Individual differences in stress-related biological measures have also been associated with temperament. The hypothalamus is the neural center for stress responses. When a stress response is triggered, the hypothalamus sends signals to the pituitary gland and the adrenal medulla. Short-term responses are produced via the sympathomedullary pathway (SAM) while long-term stress is regulated by the hypothalamic-pituitary-adrenocortical (HPA) axis. Much of the research linking temperament to behavior problems have investigated the end product of the HPA system—cortisol. Whereas baseline levels of cortisol support the individual's ability to organize and complete fight/flight responses to threatening situations, the regulation of behavioral and sympathetic reactivity is facilitated by smaller elevations of cortisol (Gunnar, Kryzer, Van Ryzin, & Phillips, 2011). Thus, cortisol responses to stressor/challenges serve an important role in challenging situations, and individual differences in reactivity and regulation are believed to prompt variations in these biological responses. Many researchers have hypothesized that behaviorally inhibited children have a lower threshold for distress and therefore produce more cortisol in response to challenge (Kagan, 1994; Rosen & Schulkin, 1998). In support of this hypothesis, higher levels of cortisol were found to be associated with greater shyness/fearful behavior in children (de Haan, Gunnar, Tout, Hart, & Stansbury, 1998; Kagan et al., 1987a; Watamura, Donzella, Alwin, & Gunnar, 2003). Researchers have also examined changes in children's cortisol over time while also considering the child's temperament. For example, Blair, Peters, and Granger (2004) reported that inhibited preschool-age children showed increases in cortisol across a laboratory visit, while exuberant children showed a drop in cortisol across this time period. Temperamental differences in changes in cortisol also appear across a longer period of time. Tarullo, Mliner, and Gunnar (2011) found that cortisol decreased from fall assessments to spring assessments for highly exuberant and average children but not for inhibited children. However, the association between cortisol and temperament traits appears to depend on other factors, both internal and external. External contexts, such as parenting (Nachmias, Gunnar, Mangelsdorf, Parritz, & Buss, 1996) and day care (de Haan et al., 1998) can affect the degree to which temperament and cortisol responsivity is linked. Indeed, individual characteristics, such as social competence, peer rejection, and emotion dysregulation, are related to cortisol levels when measured in child care contexts (Dettling, Parker, Lane, Sebanc, & Gunnar, 2000; Gunnar, Tout, de Haan, Pierce, & Stansbury, 1997; Gunnar et al., 2003). The association between temperament and cortisol may also depend on other behaviors of the child. Gunnar and colleagues (2003) reported that exuberant children who were more aggressive and experienced greater peer rejection had higher cortisol levels, whereas exuberant children who did not show these maladaptive social behaviors had lower cortisol levels than children lower in exuberance. Tarullo and colleagues (2011) also found evidence that the social experiences predictive of changes in children's cortisol over time (across an academic year) was different for inhibited and exuberant children. Links between HPA activity and problem behaviors have also been well established such that high levels of cortisol have been found in children with either type of internalizing behavior, anxiety or depression (Gunnar, 2001; Hastings, Fortier, Utendale, Simard, & Robaey, 2009; Lopez-Duran, Kovacs, & George, 2009), whereas children with externalizing problems or

disruptive behavior disorders have lower cortisol levels before and after challenging events (Fairchild, Van Goozen, Stollery, & Goodyer, 2008; Hastings et al., 2009; van Goozen, Fairchild, Snoek, & Harold, 2007). A few studies have examined the role that both temperament and cortisol reactivity have in predicting problem behaviors; the findings suggest that gender plays a significant role in this developmental process. Hastings and colleagues (2011) found that inhibition and prolonged elevations of cortisol were linked to internalizing behaviors, but only for girls. In another study, chronic inhibition, which was associated with adolescent anxiety disorder, was predicted by both earlier behavioral inhibition and high levels of cortisol, but only for girls (Essex, Klein, Slattery, Goldsmith, & Kalin, 2010). Interestingly, Perez-Edgar, Schmidt, Henderson, Schulkin, and Fox (2008) found a similar relationship for boys in that high levels of cortisol and high negativity in infancy were related to internalizing behaviors and social reticence in 4-year-old boys. One end product of stimulation of the SAM is salivary alpha-amylase (sAA), a digestive protein found in saliva and responsive to psychological and physical stress. Similar to cortisol, sAA levels change in response to challenges and stressors presented to the individual but occur in less time because of its short-term flight/fight function. As such, sAA has been used as a surrogate marker of SNS activation. However, only recently have studies examined the link between sAA and children's temperament and problem behavior. Spinrad and colleagues (2009) reported that parent ratings of temperamental anger, impulsivity, and externalizing behaviors were negatively associated with sAA reactivity, whereas regulation was positively associated with sAA, especially for girls. In another study on toddlers, high levels of baseline sAA were related to observed positive affect and approach, and the highest levels were found in toddlers rated high in both positivity and approach (Fortunato, Dribin, Granger, & Buss, 2008). Because both sAA and cortisol are used to assess stress, researchers have turned to examining how they interact to affect individual differences in behavior (El-Sheikh, Erath, Buckhalt, Granger, & Mize, 2008; Gordis, Granger, Susman, & Trickett, 2006; Keller, El-Sheikh, Granger, & Buckhalt, 2012; Kivlighan & Granger, 2006). For instance, high sAA and low levels of cortisol predicted higher aggression in young adolescents, whereas high sAA and high cortisol predicted internalizing behavior problems (Gordis et al., 2006). El-Sheikh and colleagues (2008) also found that school-age children had increased risk of internalizing behavior problems when they had high levels of cortisol and high sAA. Although promising, research that has examined sAA, temperament, and developmental psychopathology is not conclusive, and we await further findings.

Genetic Measures The heritability of temperament traits is an important part of most temperament models and research. Indeed, at least one temperament model rests on evidence that temperament traits are sufficiently heritable to be included in the model (A. Buss & Plomin, 1975, 1984). Quantitative behavioral genetics research, which uses twin and adoption study designs to decompose the observed variance of a temperament trait into genetic and environmental variance, has provided strong support that many temperament traits have a genetic basis. Temperament traits

such as activity level (Saudino & Eaton, 1991; A. Wood et al., 2007), anger/frustration (Goldsmith, Buss, & Lemery, 1997; Goldsmith & Gottesman, 1981), behavioral inhibition/shyness (Eley et al., 2003; Emde et al., 1992; Smith et al., 2012), and effortful control (Gagne & Saudino, 2010; H. Goldsmith, Pollak, & Davidson, 2008) have been investigated for the genetic effects, and evidence has been found across development. Based on eight twin studies, Goldsmith et al. (2008) suggest that individual differences in children's effortful control are at least moderately heritable. Although estimates of heritability vary across studies and traits, the evidence indicates that genetic differences account for approximately 20% to 60% of the variability in temperament within the population (Saudino & Wang, 2012). Several studies have shown shared genetic influence between temperament and measures of problem behavior (Goldsmith & Lemery, 2000; Saudino & Wang, 2012). For example, observed inhibitory control is genetically linked to externalizing problems (Gagne, Saudino, & Asherson, 2011) while genetic influence explained about 94% of the association between emotionality and internalizing problems (Schmitz et al., 1999). As noted, these findings lend support to the spectrum model linking temperament to the development of psychopathology (Tackett, 2006). Quantitative genetic research also focuses on how the environment influences the expression of temperament traits. Although much of the research shows that shared genes, not shared environments, account for similarity in temperaments (e.g., Gagne, Saudino, & Cherry, 2003; Lemery-Chalfant et al., 2008; Mullineaux, Deater-Deckard, Petrill, Thompson, & DeThorne, 2009), findings indicate that shared and nonshared environmental effects may influence the expression of temperament. For example, differential treatment by parents may affect differences in siblings' expression of shyness and activity level (Saudino, Wertz, Gagne, & Chawla, 2004). Further, although rare in temperament research, evidence of shared environmental effects for temperamental positive affect (Gagne, Vendlinski, & Goldsmith, 2009; Goldsmith, Lemery, Buss, & Campos, 1999; Goldsmith et al., 1997;) has emerged. Thus, environmental differences within and between families are important in understanding the genetic bases of temperament (Saudino & Wang, 2012). Recent developments in molecular genetic techniques now make it possible to identify specific genes associated with various temperament traits with the understanding that no one gene is responsible for individual differences in reactivity and regulation. Thus far, molecular genetics research has identified several candidate genes related to temperament with the genes regulating dopamine and serotonin the most commonly studied genes. The dopamine genes, DRD2, DRD4, and DAT1, for example, have been implicated in reward sensitivity and novelty seeking. Genes regulating serotonergic functioning, such as the 5-HTT gene, are involved in the regulation of emotion and mood. A number of studies examining these candidate genes in samples of infants, toddlers, and children have found support for these functions; however, an equal number found null or opposing associations (see Saudino & Wang, 2012, for a review). Importantly, some of the genes associated with temperament are often those associated with clinical disorders, such as the relationship between DRD4 and ADHD (Faraone et al., 2005).

As noted, the relationship between these candidate genes and temperament is not monotonic; other genes likely work together to explain variations in temperament. Further, environmental or experiential influences appear to affect the expression of temperament for children possessing certain genes. For example, Sheese, Voelker, Rothbart, and Posner (2007) found that quality of parenting interacted with the 7-repeat allele of the DRD4 gene to affect children's activity level, sensation seeking, and impulsivity. In another study, the short 5-HTT allele was associated with behavioral inhibition, but only for those who experienced low maternal support (Fox et al., 2005). Along these same lines, Moffitt, Caspi, and Rutter (2006) have skillfully proposed that gene × environment effects are a major determinant of psychopathology and suggest a number of strategies to test this hypothesis. Summary Noninvasive electrophysiological and neuroimaging procedures such as electrocardiography, EEG, ERP, functional magnetic resonance imaging, and cortisol/alpha amylase assays with infants, children, and adolescents have provided insight into the biological link between temperament and psychopathology. Notably, when taken together, the findings appear to coalesce around physiological changes that characterize fearfulness, behavioral inhibition, and anxiety. High heart rate, low resting RSA, greater cortisol reactivity (particularly when paired with higher levels of sAA), right frontal hemispheric asymmetry, and enhanced amygdala activation have been consistently found in fearful children, supporting Kagan's (1994; Kagan & Fox, 2006; Kagan et al., 1984) long-held hypothesis that a low threshold for physiological arousal is a central characteristic of behavioral inhibition. Although less studied, fearless, more approach-oriented, and reward-sensitive children have been found to exhibit low heart rate, less sympathetic reactivity, and greater left frontal activation. Recently, Miskovic and Schmidt (2012) proposed a lateralized brain-body emotion model to integrate these findings and apply them to temperament, specifically fearfulness. Central to this model is the lateralization of central and autonomic function underlying approach-withdrawal whereby the right hemisphere has an excitatory role whereas the role of the left hemisphere is to generate inhibitory responses. Support for this model comes from studies in adults examining at least two biological systems, such as EEG and RSA. Few studies have been done with children, but with the availability of child-friendly, noninvasive techniques and models to test, research in temperament and psychopathology needs to advance toward considering genetic and multiple physiological levels of analysis.

Final Remarks: Summary and Future Directions Currently there is little doubt that temperament makes up a considerable part of what the children bring to bear on their own development, including the emergence of emotional and behavioral disorders (Zentner & Shiner, 2012). Whether through extremes in temperamental reactivity or in interaction with proximal and distal contexts that increase the risk of psychological maladjustment, the influence of temperament on the development of psychopathology has been grounded in strong theoretical models and supported by empirical

investigations. The recent surge in studies investigating the developmental pathways from temperament to psychopathology using sophisticated methods, multiple levels of analyses, and diverse samples has significantly improved our understanding of the processes by which a child's individual differences in reactivity and regulation directly and indirectly impacts his or her mental health outcomes. Despite this progress, there still remain methodological issues and knowledge gaps that need to be addressed. Next we summarize the findings, highlight the issues, and propose directions for future research. Existing research has revealed relatively consistent, moderate associations between children's behavior problems and a variety of temperament factors, styles, and traits; yet there is evidence that the association between some temperament traits and behavior problems is stronger than others, such as with effortful control. Further, and as might be expected, some temperament factors/styles/traits are more strongly associated with behavior problems within the externalizing realm than the internalizing realm or vice versa. For example, high levels of anger/frustration are more likely to be related to externalizing behaviors whereas greater sadness and fear reactivity are more consistently related to internalizing problems. There are, however, several caveats to this conclusion that warrant caution and indicate the need for more research. The temperament factor of negative emotionality has been shown to have a relatively strong association with both internalizing and externalizing behaviors. However, negative emotionality is composed of different forms of negative reactivity, each of which differs by underlying motivations and emotions (Barrett & Campos, 1987). Indeed, the finding that negative emotionality predicts both broad bands of behavior problems highlights this issue. The same could be said about difficult temperament, which has already been deemed a questionable construct. Taking a more fine-grained approach that examines specific temperament traits is recommended. By examining the link between lower-order traits and behavioral and emotional disorders, we can focus on the processes by which children develop psychopathology and achieve more precision and better predictability. We know, for example, that parents can distinguish their children's emotions in early infancy (Source & Emde, 1982). Thus, we might expect fear and anger reactivity to evoke differential parent responses, some of which may not be sensitive to the emotional reactivity of the child (e.g., ignoring fear, disciplining anger harshly). Overtime these interactions may lead to diagnosable emotional and behavioral disorders. We encourage researchers in developmental psychopathology to consider lower-order temperament traits and use theoretical models and supportive research to guide their choice of trait for inclusion in their models. It is well understood that the environment within which children develop has a significant and long-lasting impact on their cognitive, social, and emotional outcomes even before birth. Thus, when investigating the role of temperament in the development of psychopathology, researchers have considered carefully the conditions and contexts within which the child's temperament operates and its effect on adjustment. Not surprisingly, the majority of studies reviewed focused on the parenting environment. Overall, the research indicates that negative parenting behaviors, such as intrusiveness and harsh discipline, seem to exacerbate the maladaptive effects of some temperament traits (i.e., negative emotionality) in the development of behavior

problems. However, potentially poor outcomes can be tempered by positive parenting behaviors, such as warmth and sensitivity, in some cases to the point of developing the fewest problem behaviors among comparable children. Despite the accumulating evidence that the parenting environment in interaction with a child's temperament increases or decreases the risk for the development of psychopathology, there are a number of gaps in the research. Few studies, for example, have examined the transactional process that exists between children's temperament and aspects of the environment across early development despite the strong theoretical frameworks and statistical applications to guide and support such work. Research on other aspects of the environment is also needed. Other proximal processes, such as family stress or child maltreatment, that may be mediated or moderated by temperament to increase risk of psychopathology should be examined. Likewise, investigations of adverse contexts, such as poverty and neighborhood risk, are likely to demonstrate both the risk and resilience of temperament traits (Lengua & Wachs, 2012). Indeed, if a researcher is considering broad risk factors such as socioecomonic status, parent mental health, or trauma as they relate to the development of psychopathology, taking into account the child's temperament will determine to some degree whether these factors lead to either emotional and behavioral disorders or normal adjustment by way of the child's perceptions and responses to these environmental conditions. A review of the current research indicates that certain temperament traits/factors appear to also serve a protective role. Although there is still limited research examining this area of inquiry, one important exception is the number of studies that have demonstrated the role of effortful control in lowering highly negative children's risk of developing behavior problems, especially within the externalizing realm. This protective role is not surprising, given the central role of self-regulation in all developmental domains. Likewise, recent research on the temperament trait of positive emotionality suggests that under certain conditions (e.g., without accompanying disruptive behaviors or when paired with low negative emotionality), greater positive emotionality can protect against both internalizing and externalizing behavior. Although these data may be seem obvious, given the number of social, cognitive, and physical outcomes associated with this trait, a growing number of studies have linked high-intensity positive affect with problem behaviors, particularly as part of the exuberant type. Continued study of temperament × temperament interactions with effortful control and positive emotionality may elucidate what combinations of characteristics put a child at risk for problem behaviors. For example, in exuberant children, do high levels of positive emotionality combined with impulsivity in particular contexts increase risk of behavior problems, such as when presented with highly novel, exciting stimuli? Or do high levels of effortful control counteract the risk conferred on children characterized by high-intensity pleasure? There is some indication that emotion regulation is important for exuberant children's behavioral adjustment (Stifter et al., 2008), but more research is needed to examine the regulation of positive affect. The work of Klein and colleagues (Dougherty et al., 2010; Klein et al., 2011) has also implicated low levels of positive emotionality in childhood depression. All children have the capacity to feel positive emotions. Future studies might examine the conditions under which these children feel and express positive affect and whether, for example, parent up-

regulation of the child's positive emotion reduces the risk of depression. Given that the examination of the link between positive emotionality and developmental psychopathology is a relatively new area of research and the findings are mixed, more studies are needed to investigate how variations in positive affect intensity, the contexts in which positive emotions are elicited, and the underlying motivations for the experience of positive emotions contribute to developmental psychopathology. More advanced statistical modeling supports new innovative methods for examining the association between temperament and the development of psychopathology, which we have highlighted in our review of the literature. For example, studies taking a person-oriented approach have used latent profile analysis to create children's temperamental styles. This technique has advanced the grouping of children based on similar characteristics from the more simplistic extreme group approach and the somewhat problematic cluster analysis, which relies on more arbitrary criteria for classification (Vermunt & Magidson, 2002). Latent growth curve modeling has been applied to longitudinal data to examine how temperament affects trajectories of behavior problems over time. This method may be useful in testing the pathoplasty/exacerbation model which proposes that personality will affect the course, severity, and presentation of a disorder. According to Tacket (2006), this model has not been tested due to the methodological challenges of measuring personality before the emergence of a disorder. If we accept that temperament forms the basis of personality (Caspi & Shiner, 2012) and that a number of clinical disorders presented in adolescence are predictive of adult psychopathology, then the field of developmental psychopathology is prepared conceptually and methodologically to take up this challenge. Researchers are increasingly implementing multiple measures in their studies of temperament. Importantly, observational and biological measures are becoming more widely incorporated into studies investigating the association between temperament and developmental psychopathology. Yet there is still room for advancement. We strongly encourage future research to consider multiple indicators of temperament across multiple levels of measurement (biological, parent report, behavioral) in single studies. A number of advanced multivariate statistics, such as structural equation modeling, can be used to address the use of multilevel, multi-informant designs. Although improved statistical analyses encourage the use of multiple methods, much of the current research continues to rely on parent ratings of both children's temperament as well as their problem behavior. The reliance on parent report for both measures of temperament and behavioral problems introduces error into the results by way of parental bias. Although there are many advantages to using parent report, research has clearly demonstrated that other factors, particularly related to the reporter, influence parent ratings of temperament. Controlling for these variables is important but not considered in a number of the studies reviewed. Future research needs to counteract the biases in parent ratings by obtaining reports from others close to the child, such as another parent or child care worker, by including observational measures of temperament, or by obtaining ratings of problem behavior from a number of informants.

Further, several studies have examined the overlap among the instruments measuring both temperament and behavior problems and found that after removing similar items, the relationship between the two constructs was attenuated but remained significant. Research following this discovery, however, has not followed suit, with the majority of studies neglecting to remove the confounding items, thus inflating the relationship between temperament and problem behaviors. For researchers who rely solely on parent report of both temperament and behavior problems, such as in studies using large community samples, removing the confounding items is highly recommended as it is essential to the integrity of the findings. Another methodological issue that limits interpretation is that the majority of the studies reviewed were conducted with community samples. In many cases, the results linking temperament with problem behavior did not have sufficient numbers of children meeting the criteria for diagnosable externalizing or internalizing clinical disorders. Thus, it is not clear to what extent temperament, even at extreme levels, influences the development of psychopathology. Future research needs to take the next step toward conducting studies that either target children at risk, such as children of depressed mothers (e.g., S. Goodman et al., 2011), or children living in violent homes (e.g., Towe-Goodman, Stifter, Coccia, & Cox, 2011), to increase the chances of finding children with clinical disorders, or use large and diverse community samples that produce sufficient numbers of children with diagnosable levels of behavioral or emotional problems to ensure power to test hypotheses linking temperament to psychopathology.

Translational Implications We end this chapter with a brief discussion of the translational implications of temperament for the field of developmental psychopathology. Temperament research emerged out of a clinical tradition; however, the application of what we have come to learn about temperament and behavioral adjustment to practice has not advanced. Only a handful of interventions focused on temperament have been conducted with the aim of reducing behavioral and emotional disorders. However, many of the methodological issues found in the studies reviewed in this chapter are also evident in these programs (McClowry, Snow, Tamis-Lemonda, & Rodrigues, 2010; Rapee, Kennedy, Ingram, Edwards, & Sweeney, 2010). Nonetheless, the research findings from the studies we reviewed in this chapter are informative in terms of understanding the developmental pathways toward clinical disorders as well as informing preventive interventions. One way to conceive of the role of child temperament in the development of psychopathology is as a risk or protective factor. Risk is a probabilistic concept that defines the likelihood that certain individual or contextual characteristics will contribute to the development of a negative outcome. Protective factors are those characteristics that reduce the impact of risk (Lengua & Wachs, 2012). The developmental impact of a risk factor will depend on a number of circumstances, including whether the child is exposed to other risks, the context within which the risk occurs, and whether there are protections against the risk. Thus, certain temperament

traits can be either a risk (e.g., negative emotionality) or a protective factor (e.g., effortful control), as the current review has highlighted. In addition, a temperament trait or style may be a risk factor in some contexts or during some developmental periods but not others. For example, anger reactivity, particularly during the preschool period, is associated with externalizing behaviors, but anger in infancy is adaptive in that it communicates the child's emotional state and elicits caregivers to change the situation (Stifter & Fox, 1990). Treating temperament traits or styles as risk or protective factors implies that it has a translational function. The findings of the most current research in temperament and psychopathology indicate that extremes in the temperament traits that define behavioral inhibition may be a risk factor for internalizing problems generally and anxiety specifically, supporting the spectrum model. Importantly, the data suggest that it is not just the presence of extreme behavioral inhibition in early childhood that denotes risk likely because many behaviorally inhibited children tend to move out of that class due to the development of regulatory strategies or other moderating influences. Rather, if a child remains behaviorally inhibited across early childhood, he or she will most likely develop internalizing problems or anxiety. This conclusion is supported by research on the biological indices underlying behavioral inhibition and anxiety, such as right hemispheric activation, as well as the processes through which such children may develop anxiety disorders (e.g., attention to threat). If stable behavioral inhibition is a risk factor, then mental health programs aimed at preventing the development of internalizing disorders in children might select those characterized by high behavioral inhibition over time and focus their interventions on processes that decrease the risk of psychopathology, such as parenting or self-regulation. A handful of interventions with behaviorally inhibited children have been conducted aiming the intervention on parenting behaviors hypothesized to attenuate the risk for anxiety (LaFreniere & Capuano, 1997; Rapee et al., 2010). Although somewhat successful in the short term, the children were classified based on a one-time assessment of behavioral inhibition as preschoolers. Drawing from the research reviewed, the interventions may have been more successful if the researchers had further limited their intervention group to those children who were stable in behavioral inhibition from earlier in childhood. Taking an extreme group approach to understanding risk for developmental psychopathology may be especially informative for practice with the behaviorally inhibited type; however, most research on other temperament traits and styles supports the vulnerability model's proposed pathways to developmental psychopathology. The vulnerability model differs from the spectrum model in that it makes a distinction between temperament and psychopathology. And while it proposes that individual differences that make up temperament and personality are risk factors for particular clinical disorders, it includes the potential for the interaction with environmental factors to increase risk. We would also add that based on the research reviewed here, combinations of temperament traits found in temperament × temperament interactions increase the risk of or provide protection against the development of psychopathology. Understandably, parent behavior, personality, and mental health have been the focus on much of the research testing the vulnerability model. Parents are particularly important in providing the context within which developmental tasks are accomplished. Likewise, depending on the type

of behavior they exhibit in response to their child's emotions and behaviors, parents may influence whether the child's developmental trajectories are positive or negative. Negative parent behaviors, such as harsh discipline or poor parent mental health, are risk factors for child behavior problems, but when those behaviors are combined with children who are high on negative emotionality and more specifically anger reactivity, which are also risk factors, then the likelihood of the child being viewed as problematic increases. Many preventive interventions aimed at reducing externalizing problems select parents and children who are identified as “at risk” due to such factors as parent psychopathology, income level, or abusive history. Understanding that temperament may add to such risks via how children perceive, react to, and subsequently evoke parental responses could be helpful in creating more selective interventions. For example, low-income parents who are easily frustrated may respond to their highly active, impulsive, surgent child with negative control. Such parenting behavior has been shown to be ineffective with surgent children, increasing their risk of developing externalizing behaviors. Selective interventions that target high-risk parents (low income, easily frustrated) with surgent children may be more effective than an intervention program that targets just the family/parent characteristics. Finally, we consider the implications of temperament × temperament interactions in the development of psychopathology. The research has clearly demonstrated that effortful control which emerges at the end of the first year of life through the preschool years is important to the regulation of temperamental reactivity. Thus, children who have poor effortful control will exhibit high levels of negative or even positive reactivity that can lead to behavior problems. These findings support current preventive interventions that focus on improving self-regulation or effortful control. However, the evidence also suggests that effortful control may not be a protective factor for the behaviorally inhibited child, demonstrating the importance of considering the child's temperament when designing preventive interventions. Positive emotionality may also act as a protective factor for other temperament traits. The findings indicate that high positive affect may attenuate the risk of negative emotionality and high activity level. For example, exuberant children, perhaps through their high active, impulsive behavior, may be at risk for problem behaviors, but the negative effects of these behaviors may be counteracted by their positive emotionality and approach orientation. According to Fredrickson's (2001) Broaden and Build theory, positive affect builds social, cognitive, and physical resources that individuals can draw on to cope with aversive situations. Exuberant children, therefore, may have protective factors within their own group of characteristics. Positive emotionality has a significant impact on all developmental domains, making it an excellent target for interventions, in particular universal interventions that are geared toward all families regardless of risk. Up-regulating positive emotions and learning to use them to cope with aversive events (positive reappraisal) may prove to be very effective in preventing the development of behavior problems.

Conclusion Historically, temperament has been proposed to have a significant influence on the

development of psychopathology. Recent improvements in the conceptualization and measurement of temperament and the progress made by studies examining the biological and genetic underpinnings of temperament have enabled researchers to test several proposed models of psychopathology. Overall, the findings are promising and support the notion that, through their individual differences in reactivity and regulation, children contribute to their own development, including their risk for emotional and behavioral disorders. Although the role of temperament in child development is no longer debatable, the degree to which temperament negatively and positively affects maladaptive outcomes and the explanatory mechanisms through which it makes its impact still requires further study.

References Achenbach, T. (2006). As others see us. Clinical and research implications of cross-informant correlations for psychopathology. Current Directions in Psychological Science, 15(2), 94–98. doi: 10.1111/j.0963–7214.2006.00414.x Achenbach, T., & Edelbrock, C. (1978). The classification of child psychopathology: A review and analysis of empirical efforts. Psychological Bulletin, 85, 1275–1301. Achenbach, T., & Edelbrock, C. (1983). Manual for the Child Behavior Checklist and Revised Child Behavior Profile. Department of Psychiatry of the University of Vermont. Achenbach, T., Edelbrock, C., & Howell, C. (1987). Empirically based assessment of the behavioral/emotional problems of 2- and 3-year-old children. Journal of Abnormal Child Psychology, 15, 629–650. Achenbach, T., McConaughy, S., & Howell, C. (1987). Child/adolescent behavioral and emotional problems: Implications of cross-informant correlations for situational specificity. Psychological Bulletin, 101(2), 213–232. doi: 10.1037/0033–2909.101.2.213 Ahadi, S., Rothbart, M., & Ye, R. (1993). Children's temperament in the US and China: Similarities and differences. European Journal of Personality, 7(5), 359–378. Aksan, N., & Kochanska, G. (2004). Heterogeneity of joy in infancy. Infancy, 6(1), 79–94. doi: 10.1207/s15327078in0601_4 Angold, A., Costello, J., & Erkanli, A. (1999). Comorbidity. Journal of Child Psychology and Psychiatry, 40, 57–87. Auerbach, J., Berger, A., Atzaba-Poria, N., Arbelle, S., Cypin, N., Friedman, A., & Landau, R. (2008). Temperament at 7, 12, and 25 months in children at familial risk for ADHD. Infant and Child Development, 17, 321–338. doi: 10.1002/icd.579 Bakermans-Kranenburg, M., & van IJzensoorn, M. (2006). Gene-environment interaction of the dopamine D4 receptor (DRD4) and observed maternal insensitivity predicting externalizing behavior in preshoolers. Developmental Psychobiology, 48, 406–409.

Barkley, R. (1997). Behavioral inhibition, sustained attention, and executive functions: Constructing a unifying theory of ADHD. Psychological Bulletin, 121, 65–94. doi: 10.1037/0033–2909.121.1.65 Barrett, K. C., & Campos, J. J. (1987). Perspectives on emotional development II: A functionalist approach to emotions. In J. D. Osofsky (Ed.), Handbook of infant development (2nd ed., pp. 555–578). New York, NY: Wiley. Bates, J. (1980). The concept of difficult temperament. Merrill-Palmer Quarterly, 26, 299– 319. Bates, J. (1989). Concepts and measures of temperament. In G. Kohnstamm, J. Bates, & M. Rothbart (Eds.), Temperament in childhood (pp. 3–26). Oxford, UK: Wiley. Bates, J., & Bayles, K. (1984). Objective and subjective components in mothers' perceptions of their children from age 6 months to 3 years. Merrill-Palmer Quarterly, 30, 111–130. Bates, J., Bayles, K., Bennett, D., Ridge, B., & Brown, M. (1991). Origins of externalizing behavior problems at eight years of age. In D. Pepler & K. Rubin (Eds.), The development and treatment of childhood aggression (pp. 93–120). Hillsdale, NJ: Erlbaum. Bates, J., Maslin, C., & Frankel, K. (1985). Attachment security, mother–child interaction, and temperament as predictors of behavior-problem ratings at age three years. Monographs of the Society for Research in Child Development, 50(1–2), 167–193. doi: 10.2307/3333832 Bates, J., & Pettit, G. (2007). Temperament, parenting, and socialization. In J. E. Grusec & P. D. Hastings (Eds.), Handbook of socialization: Theory and research (pp. 153–177). New York, NY: Guilford Press. Bayley, N. (1969). Bayley Scales of Infant Development. New York, NY: Psychological Corporation. Beauchaine, T. (2001). Vagal tone, development, and Gray's motivational theory: Toward an integrated model of autonomic nervous system functioning in psychopathology. Development and Psychopathology, 13, 183–214. Beauchaine, T., Gatzke-Kopp, L., & Mead, H. (2007). Polyvagal Tteory and developmental psychopathology: Emotion dysregulation and conduct problems from preschool to adolescence. Biological Psychology, 74(2), 174–184. doi: 10.1016/j.biopsycho.2005.08.008 Beauchaine, T., Katkin, E., Strassberg, Z., & Snarr, J. (2001). Disinhibitory psychopathology in male adolescents: Discriminating conduct disorder from attention-deficit/hyperactivity disorder through concurrent assessment of multiple autonomic states. Journal of Abnormal Psychology, 110(4), 610–624. doi: 10.1037/0021–843X.110.4.610 Bell, R. Q. (1968). A reinterpretation of the direction of effects in studies of socialization. Psychological Review, 75(2), 81–95.

Belsky, J. (1997). Theory testing, effect-size evaluation, and differential susceptibility to rearing influence: The case of mothering and attachment. Child Development, 68(4), 598–600. Belsky, J., Bakermans-Kranenburg, M., & van IJzendoorn, M. (2007). For better and for worse: Differential susceptibility to environmental influences. Current Directions in Psychological Science, 16, 300–304. doi: 10.1111/j.1467–8721.2007.00525.x Belsky, J., Friedman, S., & Hsieh, K. (2001). Testing a core emotion-regulation prediction: Does early attentional persistence moderate the effect of infant negative emotionality on later development? Child Development, 72(123–133). doi: 10.1111/1467–8624.00269 Belsky, J., Hsieh, K., & Crnic, K. (1998). Mothering, fathering, and infant negativity as antecedents of boys' externalizing problems and inhibition at age 3 years: Differential susceptibility to rearing experience? Development and Psychopathology, 10(301–319). doi: 10.1017/S095457949800162X Belsky, J., Pasco Fearon, R., & Bell, B. (2007). Parenting, attention and externalizing problems: Testing mediation longitudinally, repeatedly and reciprocally. Journal of Child Psychology and Psychiatry, 48, 1233–1242. doi: 10.1111/j.1469–7610.2007.01807.x Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to environmental influences. Psychological Bulletin, 135(6), 885–908. Bergman, L. R., & Magnusson, D. (1997). A person-oriented approach in research on developmental psychopathology. Development and Psychopathology, 9(2), 291–319. doi: 10.1017/S095457949700206X Berntson, G., & Cacioppo, J. (2004). Heart rate variability: Stress and psychiatric conditions. In M. Malik & A. Camm (Eds.), Dynamic electrocardiography (pp. 57–64). New York, NY: Blackwell Futura. Berntson, G., Cacioppo, J., Binkley, P., Uchino, B., Quigley, K., & Fieldstone, A. (1994). Autonomic cardiac control. III. Psychological stress and cardiac response in autonomic space as revealed by pharmacological blockades. Psychophysiology, 31(6), 599–608. doi: 10.1111/j.1469–8986.1994.tb02352.x Berntson, G., Cacioppo, J., & Quigley, K. (1991). Autonomic determinism: The modes of autonomic control, the doctrine of autonomic space, and the laws of autonomic constraint. Psychological Review, 98, 459–487. Biederman, J., Hirshfeld-Becker, D., Rosenbaum, J., Herot, C., Friedman, D., Snidman, N.,… Faraone, S. (2001). Further evidence of association between behavioral inhibition and social anxiety in children. American Journal of Psychiatry, 158, 1673–1679. Biederman, J., Rosenbaum, J., Bolduc-Murphy, E., & Faraone, S. (1993). A 3-year follow-up of children with and without behavioral inhibition. Journal of the American Academy of Child and Adolescent Psychiatry, 32, 814–821. doi: 10.1097/00004583–199307000–00016

Biederman, J., Rosenbaum, J., Hirshfeld, D. R., Farone, S., Bolduc, E., & Gersten, M. (1990). Psychiatric correlates of behavioral inhibition in young children of parents with and without psychiatric disorders. Archives of General Psychiatry, 47, 21–26. Bishop, G., Spence, S. H., & McDonald, C. (2003). Can parents and teachers provide a reliable and valid report of behavioral inhibition? Child Development, 74(6), 1899–1917. doi: 10.1046/j.1467–8624.2003.00645.x Blair, C., & Peters, R. (2003). Physiological and neurocognitive correlates of adaptive behavior in preschool among children in Head Start. Developmental Neuropsychology, 24(1), 479–497. doi: 10.1207/S15326942DN2401_04 Blair, C., Peters, R., & Granger, D. (2004). Physiological and neuropsychological correlates of approach/withdrawal tendencies in preschool: Further examination of the behavioral inhibition system/behavioral activation system scales for young children. Developmental Psychobiology, 45(3), 113–124. doi: 10.1002/dev.20022 Blair, C., Zelazo, P., & Greenberg, M. (2005). The measurement of executive function in early childhood. Developmental Neuropsychology, 28(2), 561–571. doi: 10.1207/s15326942dn2802_1 Blandon, A., Calkins, S., Keane, S., & O'Brien, M. (2010). Contributions of child's physiology and maternal behavior to children's trajectories of temperamental reactivity. Developmental Psychology, 46(5), 1089–1102. doi: 10.1037/a0020678 Block, J. (1971). Lives through time. Berkeley, CA: Bancroft. Bohlin, G., & Hagekull, B. (2009). Socio-emotional development: From infancy to young adulthood. Scandinavian Journal of Psychology, 50, 592–601. doi: 10.1111/j.1467– 9450.2009.00787.x Boyce, W., Quas, J., Alkon, A., Smider, N., Essex, M., & Kupfer, D. (2001). Autonomic reactivity and psychopathy in middle childhood. British Journal of Psychiatry, 179, 144–150. Boyce, W. T., & Ellis, B. J. (2005). Biological sensitivity to context: I. An evolutionarydevelopmental theory of the origins and functions of stress reactivity. Development and Psychopathology, 17(2), 271–301. Branje, S., van Lieshout, C., & van Aken, M. (2005). Relations between agreeableness and perceived support in family relationships: Why nice people are not always supportive. International Journal of Behavioral Development, 29, 120–128. doi: 10.1080/01650250444000441 Braungart, J., Plomin, R., DeFries, J., & Fulker, D. (1992). Genetic influence on tester-rated infant temperament as assessed by Bayley's Infant Behavior Record: Non-adoptive and adoptive siblings and twins. Developmental Psychology, 28, 40–47.

Braungart-Reiker, J., & Stifter, C. (1996). Infants responses to frustrating situations: Continuity and change in reactivity and regulation. Child Development, 67, 1767–1779. Broidy, L. M., Nagin, D. S., Tremblay, R. E., Bates, J. E., Brame, B., Dodge, K. A.,…Vitaro, F. (2003). Developmental trajectories of childhood disruptive behaviors and adolescent delinquency: A six-site, cross-national study. Developmental Psychology, 39(2), 222–245. doi: 10.1037/0012–1649.39.2.222 Buss, A., & Plomin, R. (1975). A temperament theory of personality development. New York, NY: Wiley. Buss, A., & Plomin, R. (1984). Temperament: Early developing personality traits. Hillsdale, NJ: Erlbaum. Buss, K. (2011). Which fearful toddlers should we worry about? Context, fear regulation, and anxiety risk. Developmental Psychology, 47, 804–819. doi: 10.1037/a0023227 Buss, K., Davidson, R., Kalin, N., & Goldsmith, H. (2004). Context-specific freezing and associated physiological reactivity as a dysregulated fear response. Developmental Psychology, 40, 583–594. Buss, K., Davis, E., Kiel, E., Brooker, R., Beekman, C., & Early, M. (2013). Dysregulated fear predicts social wariness and social anxiety symptoms during kindergarten. Journal of Clinical Child and Adolescent Psychology, 42, 603–616. Buss, K., & Goldsmith, H. (2000). Manual and normative data for the Laboratory Temperament Assessment Battery—Toddler Version. Madison, WI: University of Wisconsin, Department of Psychology. Calkins, S. (1997). Cardiac vagal tone indices of temperamental reactivity and behavioral regulation in young children. Developmental Psychobiology, 31(2), 125–135. doi: 10.1002/(SICI)1098–2302(199709)31:23.0.CO;2-M Calkins, S., Blandon, A., Williford, A., & Keane, S. (2007). Biological, behavioral, and relational levels of resilience in the context of risk for early childhood behavior problems. Development and Psychopathology, 19(03), 675–700. doi: 10.1017/S095457940700034X Calkins, S., & Dedmon, S. (2000). Physiological and behavioral regulation in two-year-old children with aggressive/destructive behavior problems. Journal of Abnormal Child Psychology, 28(2), 103–118. doi: 10.1023/A:1005112912906 Calkins, S., Dedmon, S., Gill, K., Lomax, L., & Johnson, L. (2002). Frustration in infancy: Implications for emotion regulation, physiological processes, and temperament. Infancy, 3, 175–197. Calkins, S., & Fox, N. (2002). Self-regulatory processes in early personality development: A multilevel approach to the study of childhood social withdrawal and aggression. Development

and Psychopathology, 14, 477–498. doi: 10.1017/S095457940200305X Calkins, S., Fox, N., & Marshall, T. (1996). Behavioral and physiological antecedents of inhibited and uninhibited behavior. Child Development, 67, 523–540. doi: 10.2307/1131830 Calkins, S., Graziano, P., & Keane, S. (2007). Cardiac vagal regulation differentiates among children at risk for behavior problems. Biological Psychology, 74(2), 144–153. doi: 10.1016/j.biopsycho.2006.09.005 Calkins, S., & Keane, S. (2004). Cardiac vagal regulation across the preschool period: Stability, continuity, and implications for childhood adjustment. Developmental Psychobiology, 45, 101–112. doi: 10.1002/dev.20020 Cannon, W. B. (1929). Bodily changes in pain, hunger, fear and rage. Oxford, UK: Appleton. Capaldi, D., & Rothbart, M. (1992). Development and validation of an early adolescent temperament measure. Journal of Early Adolescence, 12(2), 153–173. Carey, W. (1970). A simplified method for measuring infant temperament. Journal of Pediatrics, 77(2), 188–194. Carver, C. (2004). Negative affects deriving from the behavioral approach system. Emotion, 4, 3–22. Carver, C., & Harmon-Jones, E. (2009). Anger is an approach-related affect: Evidence and implications. Psychological Bulletin, 135(2), 183–204. doi: 10.1037/a0013965 Carver, C., & Scheier, M. (1998). On the self-regulation of behavior. New York, NY: Cambridge University Press. Caspi, A. (1998). Personality development across the life course. In Editor in Chief, W. Damon; Volume Editor, N. Eisenberg.N. Eisenberg, Handbook of child psychology: Social, emotional, and personality development. (5th ed., Vol. 3, pp. 311–388). New York, NY: Wiley. Caspi, A., Henry, B., McGee, R., Moffit, T., & Silva, P. (1995). Temperamental origins of child and adolescent behavior problems: From age three to eighteen. Child Development, 66, 55–68. Caspi, A., Moffitt, T., Newman, D., & Silva, P. (1998). Behavioral observations at age 3 years predict adult psychiatric disorders: Longitudinal evidence from a birth cohort. In M. E. Hertzig & E. A. Farber (Eds.), Annual progress in child psychiatry and child development (pp. 319– 331). Philadelphia, PA: Brunner/Mazel. Caspi, A., & Shiner, R. (2006). Personality development. In N. Eisenberg (Ed.), Handbook of child psychology: Social, emotional, and personality Development (6th ed., Vol. 3, pp. 300– 365). Hoboken, NJ: Wiley.

Caspi, A., & Shiner, R. (2012). Temperament and the development of personality traits, adaptations, and narratives. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 497–516). New York, NY: Guilford Press. Caspi, A., & Silva, P. A. (1995). Temperamental qualities at age three predict personality traits in young adulthood: Longitudinal evidence from a birth cohort. Child Development, 66, 486–498. Chronis-Tuscano, A., Degnan, K. A., Pine, D. S., Perez-Edgar, K., Henderson, H. A., Diaz, Y., …Fox, N. A. (2009). Stable early maternal report of behavioral inhibition predicts lifetime social anxiety disorder in adolescence. Journal of the American Academy of Child and Adolescent Psychiatry, 48(9), 928–935. Cicchetti, D. (1984). The emergence of developmental psychopathology. Child Development, 55, 1–7. doi: 10.2307/1129830 Cicchetti, D. (1990). A historical perspective on the discipline of developmental psychopathology. In J. E. Rolf, A. Masten, D. Cicchetti, K. Nuechterlein, & S. Weintraub (Eds.), Risk and protective factors in the development of psychopathology (pp. 2–28). New York, NY: Cambridge University Press. Cicchetti, D. (1993). Developmental psychopathology: Reactions, reflections, projections. Developmental Review, 13, 471–502. Cicchetti, D., & Cohen, D. (1995). Developmental psychopathology. New York, NY: Wiley. Cicchetti, D., & Cohen, D. (2006). Development and psychopathology. In D. Cicchetti & D. Cohen (Eds.), Developmental psychopathology (Vol. 1, pp. 1–23). New York, NY: Wiley. Cicchetti, D., & Toth, S. (2009). The past achievements and future promises of developmental psychopathology: The coming of age of a discipline. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 50(1–2), 16–25. doi: 10.1111/j.1469–7610.2008.01979.x Cipriano, E., & Stifter, C. (2010). Predicting preschool effortful control from toddler temperament and parenting behavior. Journal of Applied Developmental Psychology, 31(3), 221–230. Clark, L. (2005). Temperament as a unifying basis for personality and psychopathology. Journal of Abnormal Psychology, 114(4), 505–521. doi: 10.1037/0021–843X.114.4.505 Clarke-Stewart, K., & Allhusen, V. (2002). Nonparental caregiving. In M. Bornstein (Ed.), Handbook of parenting: Vol. 3: Being and becoming a parent (2nd ed., pp. 215–252). Mahwah, NJ: Erlbaum. Cloninger, C. (2000). Biology of personality dimensions. Current Opinion in Psychiatry, 13, 611–616. Coccaro, E., McCloskey, M., Fitzgerald, D., & Phan, K. (2007). Amygdala and orbitofrontal

reactivity to social threat in individuals with impulsive aggression. Biological Psychiatry, 62, 168–178. doi: 10.1016/j.biopsych.2006.08.024 Cohn, J., & Kanade, T. (2007). Use of automated facial image analysis for measurement of emotion expression. In J. Coan & J. Allen (Eds.), Handbook of emotion elicitation and assessment (pp. 222–238). Oxford, UK: Oxford University Press. Cole, P., Zahn-Waxler, C., Fox, N., Usher, B., & Welsh, J. (1996). Individual differences in emotion regulation and behavior problems in preschool children [Press release]. Coleman, I., Wadsworth, M., Croudace, T., & Jones, P. (2007). Forty-year psychiatric outcomes following assessment for internalizing disorder in adolescence. American Journal of Psychiatry, 164, 126–133. Conway, A., & Stifter, C. (2012). Longitudinal antecedents of executive function. Child Development, 83, 1022–1036. Copeland, W., Shanahan, L., Costello, E., & Angold, A. (2009). Childhood and adolescent psychiatric disorders as predictors of young adult disorders. Archives of General Psychiatry, 66(7), 764–772. doi: 10.1001/archgenpsychiatry.2009.85 Costa, P. T., & McCrae, R. R. (1976). Age differences in personality structure: A cluster analytic approach. Journal of Gerontology, 31(5), 564–570. Crowell, S., Beauchaine, T., Gatzke-Kopp, L., Sylvers, P., Mead, H., & Chipman-Chacon, J. (2006). Autonomic correlates of attention-deficit/hyperactivity disorder and oppositional defiant disorder in preschool children. Journal of Abnormal Psychology, 115(1), 174–178. doi: 10.1037/0021–843X.115.1.174 Davidson, R. (1992). Anterior cerebral asymmetry and the nature of emotion. Brain and Cognition, 20(1), 125–151. doi: 10.1016/0278–2626(92)90065-T Davidson, R. (2000). Affective style, psychopathology, and resilience: Brain mechanisms and plasticity. American Psychologist, 55(11), 1196–1214. doi: 10.1037/0003–066X.55.11.1196 Davidson, R., & Fox, N. (1982). Asymmetrical brain activity discriminates between positive versus negative affect stimuli in human infants. Science, 218, 1235–1237. de Haan, M., Gunnar, M., Tout, K., Hart, J., & Stansbury, K. (1998). Familiar and novel contexts yield different associations between cortisol and behavior among 2-year-old children. Developmental Psychobiology, 33(1), 93–101. Deater-Deckard, K., Petrill, S., & Thompson, L. (2007). Anger/frustration, task persistence, and conduct problems in childhood: a behavioral genetic analysis. Journal of Child Psychology and Psychiatry, 48, 80–87. doi: 10.1111/j.1469–7610.2006.01653.x Deater-Deckard, K., Petrill, S., Thompson, L., & DeThorne, L. (2005). A cross-sectional behavioral genetic analysis of task persistence in the transition to middle childhood.

Developmental Science, 8(3), F21–F26. doi: 10.1111/j.1467–7687.2005.00407.x Degangi, G., Dipietro, J., Greenspan, S., & Porges, S. (1991). Psychophysiological characteristics of the regulatory disordered infant. Infant Behavior and Development, 14(1), 37–50. doi: 10.1016/0163–6383(91)90053-U Degnan, K., Almas, A., & Fox, N. (2010). Temperament and the environment in the etiology of childhood anxiety. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 51(4), 497–517. doi: 10.1111/j.1469–7610.2010.02228.x Degnan, K., Calkins, S., Keane, S., & Hill-Soderlund, A. (2008). Profiles of disruptive behavior across early childhood: Contributions of frustration reactivity, physiological regulation, and maternal behavior. Child Development, 79, 1357–1376. doi: 10.1111/j.1467– 8624.2008.01193.x Degnan, K., & Fox, N. (2007). Behavioral inhibition and anxiety disorders: Multiple levels of a resilience process. Development and Psychopathology, 19, 729–746. doi: 10.1017/S0954579407000363 Degnan, K., Hane, A., Henderson, H., Moas, O., Reeb-Sutherland, B., & Fox, N. (2011). Longitudinal stability of temperamental exuberance and social-emotional outcomes in early childhood. Developmental Psychology, 47(3), 765–780. doi: 10.1037/a0021316 Denham, S., McKinley, M., Couchoud, E., & Holt, R. (1990). Emotional and behavioral predictors of preschool peer ratings. Child Development, 61, 1145–1152. Dennis, T. (2006). Emotional self-regulation in preschoolers: The interplay of child approach reactivity, parenting, and control capacities. Developmental Psychology42, 84–97. De Pauw, S., & Mervielde, I. (2010). Temperament, personality and developmental psychopathology: A review based on the conceptual dimensions underlying childhood traits. Child Psychiatry and Human Development, 41, 313–329. doi: 10.1007/s10578–009–0171–8 De Pauw, S., Mervielde, I., & Van Leeuwen, K. (2009). How are traits related to problem behavior in preschoolers? Similarities and contrasts between temperament and personality. Journal of Abnormal Child Psychology, 37(3), 309–325. doi: 10.1007/s10802–008–9290–0 Depue, R., & Collins, P. (1999). Neurobiology of the structure of personality: Dopamine, facilitation of incentive motivation, and extraversion. Behavioral and Brain Sciences, 22, 491–569. doi: 10.1017/S0140525X99002046 Derryberry, D., & Reed, M. (1994). Temperament and attention: Orienting toward and away from positive and negative signals. Journal of Personality and Social Psychology, 66, 1128– 1139. doi: 10.1037/0022–3514.66.6.1128 Derryberry, D., & Rothbart, M. (1997). Reactive and effortful processes in the organization of temperament. Development and Psychopathology, 9, 633–652.

Dettling, A., Parker, S., Lane, S., Sebanc, A., & Gunnar, M. (2000). Quality of care and temperament determine changes in cortisol concentrations over the day for young children in childcare. Psychoneuroendocrinology, 25(8), 819–836. Diener, M., & Kim, D. (2004). Maternal and child predictors of preschool children's social competence. Journal of Applied Developmental Psychology, 25, 3–24. doi: 10.1016/j.appdev.2003.11.006 Dollar, J. (2011). Children's positive affect and behavior problems: The role of temperamental styles, parental behaviors, and the regulation of positive affect (Doctoral dissertation). The Pennsylvania State University, University Park, PA. Dollar, J., & Stifter, C. (2012). Temperamental surgency and emotion regulation as predictors of childhood social competence. Journal of Experimental Child Psychology, 112, 178–194. doi: 10.1016/j.jecp.2012.02.004 Dougherty, L., Bufferd, S., Carlson, G., Dyson, M., Olino, T., Durbin, E., & Klein, D. (2011). Preschoolers' observed temperament and psychiatric disorders assessed with a parent diagnostic interview. Journal of Clinical Child and Adolescent Psychology, 40, 295–306. doi: 10.1080/15374416.2011.546046 Dougherty, L., Klein, D., Durbin, E., Hayden, E., & Olino, T. (2010). Temperamental positive and negative emotionality and children's depressive symptoms: A longitudinal prospective study from age three to age ten. Journal of Social & Clinical Psychology, 29(4), 462–488. Dougherty, L., Tolep, M., Bufferd, S., Olino, T., Dyson, M., Traditi, J., & Klein, D. (2013). Preschool anxiety disorders: Comprehensive assessment of clinical, demographic, temperamental, familial, and life stress correlates. Journal of Clinical Child and Adolescent Psychology, 42, 577–589. Durbin, E., Klein, D., Hayden, E., Buckley, M., & Moerk, K. (2005). Temperamental emotionality in preschoolers and parental mood disorders. Journal of Abnormal Psychology, 114, 28–37. doi: 10.1037/0021–843X.114.1.28 Durbin, E., Hayden, E., Klein, D., & Olino, T. (2007). Stability of laboratory-assessed temperamental emotionality traits from ages 3 to 7. Emotion, 7(2), 388–399. doi: 10.1037/1528–3542.7.2.388 Durbin, E., & Wilson, S. (2012). Convergent validity of and bias in maternal reports of child emotion. Psychological Assessment, 24(3), 647–660. doi: 10.1037/a0026607 Dyson, M., Olino, T., Durbin, E., Goldsmith, H., & Klein, D. (2012). The structure of temperament in preschoolers: A two-stage factor analytic approach. Emotion, 12(1), 44–57. doi: 10.1037/a0025023 Ebesutani, C., Bernstein, A., J., M., Chorpita, B., & Weisz, J. (2011). The youth self report: Applicability and validity across younger and older youths. Journal of Clinical Child and

Adolescent Psychology, 40(2), 338–346. doi: 10.1080/15374416.2011.546041 Edelbrock, C., Costello, A., Dulcan, M., Conover, N., & Kala, R. (1986). Parent-child agreement on child psychiatric symptoms assessed via structured interview. Journal of Child Psychology and Psychiatry, 27(2), 181–190. doi: 10.1111/j.1469–7610.1986.tb02329.x Eisenberg, N., Cumberland, A., Spinrad, T., Fabes, R., Shepard, S., Reiser, M., & Guthrie, I. (2001). The relations of regulation and emotionality to children's externalizing and internalizing problem behavior. Child Development, 72, 1112–1134. doi: 10.1111/1467– 8624.00337 Eisenberg, N., Cumberland, A. J., & Spinrad, T. L. (1998). Parental socialization of emotion. Psychological Inquiry, 9, 241–273. Eisenberg, N., Fabes, R., Guthrie, I., Murphy, B., Maszk, P., Holmgren, R., & Suh, K. (1996). The relations of regulation and emotionality to problem behavior in elementary school children. Development and Psychopathology, 8, 141–162. Eisenberg, N., Fabes, R., & Spinrad, T. (2006). Prosocial development. In W. D. N. Eisenberg & R. M. Lerner (Eds.), Handbook of child psychology: Vol. 3, Social, emotional, and personality development (6th ed., pp. 646–718). Hoboken, NJ: Wiley. Eisenberg, N., & Morris, A. (2002). Children's emotion-related regulation. In R. V. Kail (Ed.), Advances in child development and behavior (Vol. 30, pp. 190–229). San Diego, CA: Academic Press. Eisenberg, N., Sadovsky, A., Spinrad, T., Fabes, R., Losoya, S., Valiente, C., & Shepard, S. (2005). The relations of problem behavior status to children's negative emotionality, effortful control, and impulsivity: Concurrent relations and prediction of change. Developmental Psychology, 41, 193–211. doi: 10.1037/0012–1649.41.1.193 Eisenberg, N., Shepard, S., Fabes, R., Murphy, B., & Gutherie, I. (1998). Shyness and children's emotionality, regulation, and coping: Contemporaneous, longitudinal and acrosscontext relations. Child Development, 69, 767–790. Eisenberg, N., Spinrad, T., & Eggum, N. (2010). Emotion-related self-regulation and its relation to children's maladjustment. Annual Review of Clinical Psychology, 61, 6495–6525. doi: 10.1146/annurev.clinpsy.121208.131208 Eisenberg, N., Spinrad, T., Eggum, N., Silva, K., Reiser, M., Hofer, C., & Michalik, N. (2010). Relations among maternal socialization, effortful control, and maladjustment in early childhood. Development and Psychopathology, 22, 507–525. doi: 10.1017/S0954579410000246 Eisenberg, N., Spinrad, T., Fabes, R., Reiser, M., Cumberland, A., Shepard, S.,…Murphy, B. (2004). The relations of effortful control and impulsivity to children's resiliency and adjustment. Child Development, 75, 25–46.

Eisenberg, N., Valiente, C., Spinrad, T., Cumberland, A., Liew, J., Reiser, M., & Losoya, S. (2009). Longitudinal relations of children's effortful control, impulsivity, and negative emotionality to their externalizing, internalizing, and co-occurring behavior problems. Developmental Psychology, 45, 988–1008. doi: 10.1037/a0016213 Eisenberg, N., Zhou, Q., Spinrad, T., Valiente, C., Fabes, R., & Liew, J. (2005). Relations among positive parenting, children's effortful control, and externalizing problems: A threewave longitudinal study. Child Development, 76, 1055–1071. doi: 10.1111/j.1467–8624.2005 ….00897.x Eley, T., Bolton, D., O'Connor, T., Perrin, S., Smith, P., & Plomin, R. (2003). A twin study of anxiety-related behaviours in pre-school children. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 44(7), 945–960. Ellis, B. J., Boyce, W. T., Belsky, J., Bakermans-Kranenburg, M. J., & van Ijzendoorn, M. H. (2011). Differential susceptibility to the environment: An evolutionary-neurodevelopmental theory. Development and Psychopathology, 23(1), 7–28. doi: 10.1017/S0954579410000611 Ellis, L., Rothbart, M., & Posner, M. (2004). Individual differences in executive attention predict self-regulation and adolescent psychosocial behaviors. In R. E. Dahl & L. Spear (Eds.), Adolescent brain development: Vulnerabilities and opportunities (pp. 337–340). New York, NY: New York Academy of Sciences. El-Sheikh, M., Erath, S., Buckhalt, J., Granger, D., & Mize, J. (2008). Cortisol and children's adjustment: The moderating role of sympathetic nervous system activity. Jounal of Abnormal Child Psychology, 36(4), 601–611. doi: 10.1007/s10802–007–9204–6 El-Sheikh, M., Hager, J., & Whitson, S. (2001). Exposure to interparental conflict and children's adjustment and physical health: The moderating role of vagal tone. Child Development, 72, 1617–1636. El-Sheikh, M., Kouros, C., Erath, S., Cummings, M., Keller, P., & Staton, L. (2009). Marital conflict and children's externalizing behavior: Pathways involving interactions between parasympathetic and sympathetic nervous system activity. Monographs of the Society for Reseach in Child Development, 74, vii–79. Emde, R., Plomin, R., Robinson, J., Corley, R., DeFries, J., Fulker, D.,…Zahn-Waxler, C. (1992). Temperament, emotion, and cognition at fourteen months: The MacArthur Longitudinal Twin Study. Child Development, 63(6), 1437–1455. Essex, M., Armstrong, J., Burk, L., Goldsmith, H., & Boyce, W. (2011). Biological sensitivity to context moderates the effects of the early teacher–child relationship on the development of mental health by adolescence. Development and Psychopathology, 23, 149–161. doi: 10.1017/S0954579410000702 Essex, M., Klein, M., Slattery, M., Goldsmith, H., & Kalin, N. (2010). Early risk factors and

developmental pathways to chronic high inhibition and social anxiety disorder in adolescence. American Journal of Psychiatry, 167(1), 40–46. doi: 10.1176/appi.ajp.2009.07010051 Eysenck, H. J. (1967). The biological basis of personality. Springfield, IL: Thomas. Fairchild, G., Van Goozen, S., Stollery, S., & Goodyer, I. (2008). Fear conditioning and affective modulation of the startle reflex in male adolescents with early-onset or adolescenceonset conduct disorder and healthy control subjects. Biological Psychiatry, 63(3), 279–285. doi: 10.1016/j.biopsych.2007.06.019 Faraone, S., Perlis, R., Doyle, A., Smoller, J., Goralnick, J., Holmgren, M., & Sklar, P. (2005). Molecular genetics of attention-deficit/hyperactivity disorder. Biological Psychiatry, 57, 1313–1323. Feng, X., Shaw, D., & Silk, J. (2008). Developmental trajectories of anxiety symptoms among boys across early and middle childhood. Journal of Abnormal Psychology, 117(1), 32–47. doi: 10.1037/0021–843X.117.1.32 Foley, M., McClowry, S., & Castellanos, F. (2008). The relationship between attention deficit hyperactivity disorder and child temperament. Journal of Applied Developmental Psychology, 29, 157–169. doi: 10.1016/j.appdev.2007.12.005 Fortunato, C., Dribin, A., Granger, D., & Buss, K. (2008). Salivary alpha-amylase and cortisol in toddlers: Differential relations to affective behavior. Developmental Psychobiology, 50, 807–818. Fortunato, C., Gatzke-Kopp, L., & Ram, N. (2013). Associations between respiratory sinus arrhythmia reactivity and internalizing and externalizing symptoms are emotion specific. Cognitive, Affective, & Behavioral Neuroscience, 13(2), 238–251. doi: 10.3758/s13415– 012–0136–4 Fowles, D., Kochanska, G., & Murray, K. (2000). Electrodermal activity and temperament in preschool children. Psychophysiology, 37, 777–787. Fox, N. (1989). Psychophysiological correlates of emotional reactivity during the first year of life. Developmental Psychology, 25(3), 364–372. doi: 10.1037/0012–1649.25.3.364 Fox, N. (1994). Dynamic cerebral processes underlying emotion regulation. Monographs of the Society for Research in Child Development, 59, 152–166, 250–283. doi: 10.2307/1166143 Fox, N., & Davidson, R. (1988). Patterns of brain electrical activity during facial signs of emotion in 10-month-old infants. Developmental Psychology, 24, 230–236. Fox, N., & Henderson, H. (1999). Does infancy matter? Predicting social behavior from infant temperament. Infant Behavior & Development, 22, 445–455. doi: 10.1016/S0163– 6383(00)00018–7

Fox, N., Henderson, H., Marshall, P., Nichols, K., & Ghera, M. (2005). Behavioral inhibition: Linking biology and behavior within a developmental framework. Annual Review of Psychology, 56, 235–262. doi: 10.1146/annurev.psych.55.090902.141532 Fox, N., Henderson, H., Pérez-Edgar, K., & White, L. (2008). The biology of temperament: An integrative approach. In C. A. Luciana & M. Luciana (Eds.), Handbook of developmental cognitive neuroscience (2nd ed.) (pp. 839–853). Cambridge, MA: MIT Press. Fox, N., Henderson, H., Rubin, K., Calkins, S., & Schmidt, L. (2001). Continuity and discontinuity of behavioral inhibition and exuberance: Psychophysiological and behavioral influences across the first four years of life. Child Development, 72(1), 1–21. doi: 10.1111/1467–8624.00262 Fox, N., Rubin, K., Calkins, S., Marshall, T., Coplan, R., Porges, S.,…Stewart, S. (1995). Frontal activation asymmetry and social competence at four years of age. Child Development, 66(6), 1770–1784. doi: 10.1111/j.1467–8624.1995.tb00964.x Fox, N., Schmidt, L., Calkins, S., Rubin, K., & Coplan, R. (1996). The role of frontal activation in the regulation and dysregulation of social behavior during the preschool years. Development and Psychopathology, 8(01), 89–102. doi: 10.1017/S0954579400006982 Fredrickson, B. (2001). The role of positive emotions in positive psychology. American Psychologist, 56, 216–226. Frick, P., & Morris, A. (2004). Temperament and developmental pathways to conduct problems. Journal of Clinical Child and Adolescent Psychology, 33, 54–68. doi: 10.1207/S15374424JCCP3301_6 Frick, P., Ray, J., Thornton, L., & Kahn, R. (2014). Can callous-unemotional traits enhance the understanding, diagnosis, and treatment of serious conduct problems in children and adolescents? A comprehensive review. Psychological Bulletin, 140(1), 1–57. doi: 10.1037/a0033076 Fullard, W., McDevitt, S., & Carey, W. (1984). Assessing temperament in one- to three-yearold children. Journal of Pediatric Psychology, 9, 205–217. doi: 10.1093/jpepsy/9.2.205 Gagne, J., & Saudino, K. (2010). Wait for it! A twin study of inhibitory control in early childhood. Behavior Genetics, 40(3), 327–337. Gagne, J., Saudino, K., & Asherson, P. (2011). The genetic etiology of inhibitory control and behavior problems at 24 months of age. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 52(11), 1155–1163. doi: 10.1111/j.1469–7610.2011.02420.x Gagne, J., Saudino, K., & Cherry, S. (2003). Genetic influences on temperament in early adolescence. In S. Petrill, R. Plomin, J. DeFries, & J. Hewitt (Eds.), Nature, nurture, and the transition to early adolescence (pp. 166–184). New York, NY: Oxford University Press.

Gagne, J., Van Hulle, C., Aksan, N., Essex, M., & Goldsmith, H. (2011). Deriving childhood temperament measures from emotion-eliciting behavioral episodes: Scale construction and initial validation. Psychological Assessment, 23(2), 337–353. doi: 10.1037/a0021746 Gagne, J., Vendlinski, M., & Goldsmith, H. (2009). The genetics of childhood temperament. In Y. Kim (Ed.), Handbook of behavior genetics (pp. 251–267). New York, NY: Springer. Gandour, M. J. (1989). Activity levels as a dimension of temperament in toddlers: Its relevance for the organismic specificity hypothesis. Child Development, 60(5), 1092–1098. García Coll, C., Kagan, J., & Reznick, J. S. (1984). Behavioral inhibition in young children. Child Development, 55, 1005–1019. Garon, N., Bryson, S. E., & Smith, I. M. (2008). Executive function in preschoolers: A review using an integrative framework. Psychological Bulletin, 134(1), 31–60. Gartstein, M., Putnam, S., & Rothbart, M. (2012). Etiology of preschool behavior problems: Contributions of temperament attributes in early childhood. Infant Mental Health, 33, 197– 211. doi: 10.1002/imhj.21312 Gartstein, M., & Rothbart, M. (2003). Studying infant temperament via the Revised Infant Behavior Questionnaire. Infant Behavior and Development, 26(1), 64–86. Gazelle, H. (2006). Class climate moderates peer relations and emotional adjustment in children with an early history of anxious solitude: A child × environment model. Developmental Psychology, 42, 1179–1192. doi: 10.1037/0012–1649.42.6.1179 Gentzler, A., Santucci, A., Kovacs, M., & Fox, N. (2009). Respiratory sinus arrhythmia reactivity predicts emotion regulation and depressive symptoms in at-risk and control children. Biological Psychology, 82(2), 156–163. doi: 10.1016/j.biopsycho.2009.07.002 Gilliom, M., & Shaw, D. (2004). Codevelopment of externalizing and internalizing problems in early childhood. Development and Psychopathology, 16(2), 313–333. doi: 10.1017/S0954579404044530 Gilliom, M., Shaw, D., Beck, J. E., Schonberg, M. A., & Lukon, J. L. (2002). Anger regulation in disadvantaged preschool boys: Strategies, antecedents, and the development of self-control. Developmental Psychology, 38(2), 222–235. Glied, S., & Pine, D. (2002). Consequences and correlates of adolescent depression. Archives of Pediatrics & Adolescent Medicine, 156(10), 1009–1014. doi: 10.1001/archpedi.156.10.1009 Goldsmith, H. (1996). Studying temperament via construction of the Toddler Behavior Assessment Questionnaire. Child Development, 67(1), 218–235. Goldsmith, H., Buss, K., & Lemery, K. (1997). Toddler and childhood temperament: Expanded content, stronger genetic evidence, new evidence for the importance of environment.

Developmental Psychology, 33, 891–905. Goldsmith, H., Buss, A., Plomin, R., Rothbart, M., Thomas, A., Chess, S.,…McCall, R. (1987). Roundtable: What is temperament? Four approaches. Child Development, 58, 505– 529. Goldsmith, H., & Campos, J. (1982). Toward a theory of infant temperament. In R. Emde & R. Harmon (Eds.), The development of attachment and affiliative systems (pp. 161–193). New York, NY: Plenum. Goldsmith, H., & Campos, J. (1990). The structure of temperamental fear and pleasure in infants: A psychometric perspective. Child Development, 61, 1944–1964. Goldsmith, H., & Gagne, J. (2012). Behavioral assessment of temperament. In M. Zentner & R. L. Shiner (Eds.), Handbook of temperament (pp. 209–228). New York, NY: Guilford Press. Goldsmith, H., & Gottesman, I. (1981). Origins of variation in behavioral style: a longitudinal study of temperament in young twins. Child Development, 52(1), 91–103. Goldsmith, H., I., G., & Lemery, K. (1997). Epigenetic approaches to developmental psychopathology. Development and Psychopathology, 9(02), 365–387. doi: 10.1017/S0954579497002095 Goldsmith, H., & Lemery, K. (2000). Linking temperamental fearfulness and anxiety symptoms: A behavior–genetic perspective. Biological Psychiatry, 48(12), 1199–1209. doi: 10.1016/S0006–3223(00)01003–9 Goldsmith, H., Lemery, K., Buss, K., & Campos, J. (1999). Genetic analyses of focal aspects of infant temperament. Developmental Psychology, 35(4), 972–985. Goldsmith, H., Pollak, S., & Davidson, R. (2008). Developmental neuroscience perspective on emotion regulation. Child Development Perspectives, 2, 132–140. Goldsmith, H., Reilly, J., Lemery, K., Longley, S., & Prescott, A. (1993). Preliminary manual for the Preschool Laboratory Temperament Assessment Battery, version 1.0. Madison, WI: University of Wisconsin-Madison, Department of Psychology. Goldsmith, H., Rieser-Danner, L., & Briggs, S. (1991). Evaluating convergent and discriminant validity of temperament questionnaires for preschoolers, toddlers, and infants. Developmental Psychology, 27(4), 566–579. doi: 10.1037/0012–1649.27.4.566 Goldsmith, H., & Rothbart, M. (1996). The Laboratory Temperament Assessment Battery: Prelocomotor Version. 3.0. Madison, WI. Goodman, J., Meltzer, H., & Bailey, V. (1998). The strengths and difficulties questionnaire: A pilot study on the validity of the self-report version. European Child & Adolescent Psychiatry, 7, 125–130.

Goodman, S., Rouse, M., Connel, A., Broth, M., Hall, C., & Heyward, D. (2011). Maternal depression and child psychopathology: A meta-analysis. Clinical Child and Family Psychology Review, 14, 1–27. Gordis, E., Granger, D., Susman, E., & Trickett, P. (2006). Asymmetry between salivary cortisol and alpha-amylase reactivity to stress: Relation to aggressive behavior in adolescents. Psychoneuroendocrinology, 31(8), 976–987. doi: 10.1016/j.psyneuen.2006.05.010 Gray, J. (1987). Perspectives on anxiety and impulsivity: A commentary. Journal of Research in Personality, 21, 493–509. doi: 10.1016/0092–6566(87)90036–5 Graziano, P., & Derefinko, K. (2013). Cardiac vagal control and children's adaptive functioning: A meta-analysis. Biological Psychology, 94(1), 22–37. doi: 10.1016/j.biopsycho.2013.04.011 Gunnar, M. (2001). The role of glucocorticoids in anxiety disorders: A critical analysis. In M. Vasey & M. Dadds (Eds.), The developmental psychopathology of anxiety (pp. 143–159). New York, NY: Oxford University Press. Gunnar, M., Kryzer, E., Van Ryzin, M., & Phillips, D. (2011). The import of cortisol rise in child care differs as a function of behavioral inhibition. Developmental Psychology, 47, 792– 803. Gunnar, M., Sebanc, A., Tout, K., Donzella, B., & van Dulmen, M. (2003). Peer rejection, temperament and cortisol activity in preschoolers. Developmental Psychobiology, 43, 346– 358. Gunnar, M., Tout, K., de Haan, M., Pierce, S., & Stansbury, K. (1997). Temperament, social competence, and adrenocortical activity in preschoolers. Developmental Psychobiology, 31(1), 65–85. Hagekull, B. (1994). Infant temperament and early childhood functioning: Possible relations to the five-factor model. In C. Halverson, G. Kohnstamm, & R. Martin (Eds.), The developing structure of temperament and personality from infancy to adulthood (pp. 227–240). Hillsdale, NJ: Erlbaum. Halverson, C., Havill, V., Deal, J., Baker, S., Victor, J., Pavlopoulos, V.,…Wen, L. (2003). Personality structure as derived from parental ratings of free descriptions of children: The inventory of child individual differences. Journal of Personality, 71(6), 995–1026. Hane, A., Cheah, C., Rubin, K., & Fox, N. (2008). The role of maternal behavior in the relation between shyness and social reticence in early childhood and social withdrawal in middle childhood. Social Development, 17, 795–811. doi: 10.1111/j.1467–9507.2008.00481.x Hane, A., Fox, N., Henderson, H., & Marshall, P. (2008). Behavioral reactivity and approachwithdrawal bias in infancy. Developmental Psychology, 44(5), 1491–1496.

Hardaway, C., Wilson, M., Shaw, D., & Dishion, T. (2012). Family functioning and externalizing behaviour among low-income children: Self-regulation as a mediator. Infant and Child Development, 21, 67–84. doi: 10.1002/icd.765 Hardee, J., Benson, B., Bar-Haim, Y., Moog, K., Bradley, B., Chen, G.,…Perez-Edgar, K. (2013). Patterns of neural connectivity during an attention bias task moderate associations between early childhood temperament and internalizing symptoms in young adulthood. Biological Psychiatry, 74, 273–279. Hastings, P., Fortier, I., Utendale, W., Simard, L., & Robaey, P. (2009). Adrenocortical functioning in boys with attention-deficit/hyperactivity disorder: examining subtypes of ADHD and associated comorbid conditions. Journal of Abnormal Child Psychology, 37(4), 565–578. doi: 10.1007/s10802–008–9292-y Hastings, P., Ruttle, P., Serbin, L., Mills, R., Stack, D., & Schwartzman, A. (2011). Adrenocortical responses to strangers in preschoolers: relations with parenting, temperament, and psychopathology. Developmental Psychobiology, 53(7), 694–710. doi: 10.1002/dev.20545 Hayden, E., Durbin, E., Klein, D., & Olino, T. (2010). Maternal personality influences the relationship between maternal reports and laboratory measures of child temperament. Journal of Personality Assessment, 92(6), 586–593. doi: 10.1080/00223891.2010.513308 He, J., Degnan, K. A., McDermott, J. M., Henderson, H. A., Hane, A. A., Xu, Q., & Fox, N. A. (2010). Anger and approach motivation in infancy: Relations to early childhood inhibitory control and behavior problems. Infancy, 15(3), 246–269. doi: 10.1111/j.1532– 7078.2009.00017.x Hegvik, R., McDevitt, S., & Carey, W. (1982). The middle childhood temperament questionnaire. Journal of Developmental and Behavioral Pediatrics, 3(4), 197–200. Henderson, H., Fox, N., & Rubin, K. (2001). Temperamental contributions to social behavior: The moderating roles of frontal EEG asymmetry and gender. Journal of the American Academy of Child & Adolescent Psychiatry, 40(1), 68–74. doi: 10.1097/00004583– 200101000–00018 Henderson, H., & Wachs, T. (2007). Temperament theory and the study of cognition-emotion interactions across development. Developmental Review, 27, 396–427. doi: 10.1016/j.dr.2007.06.004 Hill, A., Degnan, K., Calkins, S., & Keane, S. (2006). Profiles of externalizing behavior problems for boys and girls across preschool: The roles of emotion regulation and inattention. Developmental Psychology, 42, 913–928. doi: 10.1037/0012–1649.42.5.913 Hirshfeld, D., Rosenbaum, J., Biederman, J., Bolduc, E., Faraone, S., Snidman, N., Reznick, J. S. & Kagan, J. (1992). Stable behavioral inhibition and its association with anxiety disorder.

Journal of the American Academy of Child and Adolescent Psychiatry, 31, 103–111. Hirshfeld-Becker, D., Micco, J., Henin, A., Bloomfield, A., Biederman, J., & Rosenbaum, J. (2008). Behavioral inhibition. Depression and Anxiety, 25, 357–367. doi: 10.1002/da.20490 Hofstra, M., Van der Ende, J., & Verhulst, F. (2002). Pathways of self-reported problem behaviors from adolescence into adulthood. American Journal of Psychiatry, 159, 401–407. doi: 10.1176/appi.ajp.159.3.401 Huffman, L., Bryan, Y., del Carmen, R., Pedersen, F., Doussard-Roosevelt, J., & Porges, S. (1998). Infant temperament and cardiac vagal tone: Assessments at twelve weeks of age. Child Development, 69(3), 624–635. doi: 10.1111/j.1467–8624.1998.tb06233.x Ilott, N., Saudino, K., Wood, A., & Asherson, P. (2010). A genetic study of ADHD and activity level in infancy. Genes, Brain, and Behavior, 9(3), 296–304. doi: 10.1111/j.1601– 183X.2009.00560.x Ingram, R., & Luxton, D. (2005). Vulnerability-stress models. In B. Hankin & J. Abela (Eds.), Developmental psychopathology: A vulnerability-stress perspective (pp. 32–46). Thousand Oaks, CA: Sage. Kagan, J. (1994). Galen's prophecy. New York, NY: Westview Press. Kagan, J. (1997). Temperament and the reactions to unfamiliarity. Child Development, 68, 139–143. doi: 10.2307/1131931 Kagan, J. (1998). Biology and the child. In W. Damon & N. Eisenberg (Eds.), Handbook of child psychology. Vol. 3, Social, emotional and personality development (pp. 178–235). New York, NY: Wiley. Kagan, J. (2002). Childhood predictors of states of anxiety. Dialogues in Clinical Neuroscience, 4, 287–293. Kagan, J., Arcus, D., & Snidman, N. (1993). The idea of temperament: Where do we go from here? In R. Plomin & G. McClearn (Eds.), Nature, nurture & psychology (pp. 197–210). Washington, DC: American Psychological Association. Kagan, J., & Fox, N. (2006). Biology, culture, and temperament biases. In W. Damon, R. Lerner, & N. Eisenberg (Eds.), Handbook of child psychology (pp. 167–225). Hoboken, NJ: Wiley. Kagan, J., Reznick, S., Clarke, C., & Snidman, N. (1984). Behavioral inhibition to the unfamiliar. Child Development, 55, 2212–2225. Kagan, J., Reznick, S., & Snidman, N. (1987a). The physiology and psychology of behavioral inhibition in children. Child Development, 58, 1459–1473. Kagan, J., Reznick, S., & Snidman, N. (1987b). Temperamental variation in response to the

unfamiliar. In N. Krasnegor, E. Blass, & M. Hofer (Eds.), Perinatal development: A psychobiological perspective (pp. 421–440). San Diego, CA: Academic Press. Kagan, J., Reznick, S., Snidman, N., Gibbons, J., & Johnson, M. (1988). Childhood derivatives of inhibition and lack of inhibition to the unfamiliar. Child Development, 59, 1580–1589. doi: 10.2307/1130672 Kagan, J., & Snidman, N. (2004). The long shadow of temperament. Cambridge, MA: Belknap Press/Harvard University Press. Kagan, J., Snidman, N., McManis, M., Woodward, S., & Hardway, C. (2002). One measure, one meaning: Multiple measures, clearer meaning. Development and Psychopathology, 14, 463–475. Karreman, A., van Tuijl, C., van Aken, M., & Deković, M. (2009). Predicting young children's externalizing problems: Interactions among effortful control, parenting, and child gender. Merrill-Palmer Quarterly, 55, 111–134. doi: 10.1353/mpq.0.0020 Keenan, K., & Shaw, D. (2003). Starting at the beginning: Exploring the etiology of antisocial behavior in the first years of life. In B. Lahey, T. Moffitt & A. Caspi (Eds.), Causes of conduct disorder and juvenile delinquency (pp. 153–181). New York, NY: Guilford Press. Keenan, K., Shaw, D., Delliquadri, E., Giovannelli, J., & Walsh, B. (1998). Evidence for the continuity of early problem behaviors: Application of a developmental model. Journal of Abnormal Child Psychology, 26, 441–452. doi: 10.1023/A:1022647717926 Keller, P., El-Sheikh, M., Granger, D., & Buckhalt, J. (2012). Interactions between salivary cortisol and alpha-amylase as predictors of children's cognitive functioning and academic performance. Physiology & Behavior, 105(4), 987–995. doi: 10.1016/j.physbeh.2011.11.005 Kerr, D., Lopez, N., Olson, S., & Sameroff, A. (2004). Parental discipline and externalizing behavior problems in early childhood: The roles of moral regulation and child gender. Journal of Abnormal Child Psychology, 32, 369–383. Kertes, D., Donzella, B., Talge, N., Garvin, M., Van Ryzin, M., & Gunnar, M. (2009). Inhibited temperament and parent emotional availability differentially predict young children's cortisol responses to novel social and nonsocial events. Developmental Psychobiology, 51(7), 521– 532. doi: 10.1002/dev.20390 Kiel, E., & Buss, K. (2010). Maternal expectations for toddlers' reactions to novelty: Relations of maternal internalizing symptoms and parenting dimensions to expectations and accuracy of expectations. Parenting Science and Practice, 10, 202–218. doi: 10.1080/15295190903290816 Kiel, E., & Buss, K. (2011). Prospective relations among fearful temperament, protective parenting, and social withdrawal: The role of maternal accuracy in a moderated mediation framework. Journal of Abnormal Child Psychology, 39(7), 953–966. doi: 10.1007/s10802–

011–9516–4 Kiff, C. J., Lengua, L. J., & Zalewski, M. (2011). Nature and nurturing: Parenting in the context of child temperament. Clinical Child and Family Psychology Review, 14(3), 251–301. doi: 10.1007/s10567–011–0093–4 Kim, J., & Deater-Deckard, K. (2011). Dynamic changes in anger, externalizing and internalizing problems: Attention and regulation. Journal of Child Psychology and Psychiatry, 52, 156–166. doi: 10.1111/j.1469–7610.2010.02301.x Kim, S., & Kochanska, G. (2012). Child temperament moderates effects of parent-child mutuality on self-regulation: A relationship-based path for emotionally negative infants. Child Development, 83, 1275–1289. Kivlighan, K., & Granger, D. (2006). Salivary alpha-amylase response to competition: Relation to gender, previous experience, and attitudes. Psychoneuroendocrinology, 31(6), 703–714. doi: 10.1016/j.psyneuen.2006.01.007 Klein, D., Dyson, M., Kujawa, A., & Kotov, R. (2012). Temperament and internalizing disorders. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 541–561). New York, NY: Guilford Press. Klein, D., Kotov, R., & Bufferd, S. (2011). Personality and depression: Explanatory models and review of the evidence. Annual Review of Clinical Psychology, 7, 269–295. doi: 10.1146/annurev-clinpsy-032210–104540 Kochanska, G. (1991). Socialization and temperament in the development of guilt and conscience. Child Development, 62, 1379–1392. doi: 10.2307/1130813 Kochanska, G. (1995). Children's temperament mothers' discipline, and security of attachment: Multiple pathways to emerging internalization. Child Development, 66, 597–615. Kochanska, G. (1997). Multiple pathways to conscience for children with different temperaments: From toddlerhood to age 5. Developmental Psychology, 33, 228–240. Kochanska, G., Aksan, N., & Joy, M. (2007). Children's fearfulness as a moderator of parenting in early socialization: Two longitudinal studies. Developmental Psychology, 43, 222–237. doi: 10.1037/0012–1649.43.1.222 Kochanska, G., Aksan, N., Penney, S., & Doobay, A. (2007). Early positive emotionality as a heterogenous trait: Implications for children's self-regulation. Journal of Personality and Social Psychology, 93(6), 1054–1066. doi: 10.1037/0022–3514.93.6.1054 Kochanska, G., Coy, K., & Murray, K. (2001). The development of self-regulation in the first four years of life. Child Development, 72, 1091–1111. Kochanska, G., Friesenborg, A., Lange, L., & Martel, M. (2004). Parents' personality and infants' temperament as contributors to their emerging relationship. Journal of Personality and

Social Psychology, 86, 744–759. doi: 10.1037/0022–3514.86.5.744 Kochanska, G., Gross, J., Lin, M., & Nicholas, K. (2002). Guilt in young children: Development, determinants, and relations with a broader system of standards. Child Development, 73, 461–482. Kochanska, G., & Kim, S. (2013). Difficult temperament moderates links between maternal responsiveness and children's compliance and behavior problems in low-income families. Journal of Child Psychology and Psychiatry, 54, 323–332. doi: 10.1111/jcpp.12002 Kochanska, G., & Knaack, A. (2003). Effortful control as a personality characteristic of young children: Antecedents, correlates, and consequences. Journal of Personality, 71(6), 1087– 1112. Kochanska, G., Murray, K., & Harlan, E. (2000). Effortful control in early childhood: Continuity and change antecedents, and implications for social development. Developmental Psychology, 18, 199–214. Kohnstamm, G. A., Mervielde, I., Besevegis, E., & Halverson, C. F. (1995). Tracing the Big Five in parents' free descriptions of their children. European Journal of Personality, 9(4), 283–304. doi: 10.1002/per.2410090405 Kolko, D., & Kazdin, A. (1993). Emotional/behavioral problems in clinic and nonclinic children: Correspondence among child, parent and teacher reports. Journal of Child Psychology and Psychiatry, 34(6), 991–1006. doi: 10.1111/j.1469–7610.1993.tb01103.x Kopp, C. B. (1982). Antecedents of self-regulation: A developmental perspective. Developmental Psychology, 18, 199–214. Krueger, R., & Tackett, J. (2003). Personality and psychopathology: Working toward the bigger picture. Journal of Personality Disorders, 17(2), 109–128. doi: 10.1521/pedi.17.2.109.23986 LaFreniere, P., & Capuano, F. (1997). Preventative intervention as a means of clarifying direction of effects in socialization: Anxious-withdrawn preschoolers case. Development and Psychopathology, 9, 551–564. Lahat, A., Degnan, K., White, L., McDermott, J., Henderson, H., Lejuez, C., & Fox, N. (2013). Temperamental exuberance and executive function predict propensity for risk taking in childhood—CORRIGENDUM. Development and Psychopathology, 25(Special Issue 01), 275–275. doi: 10.1017/S0954579412001022 Laursen, B., & Hoff, E. (2006). Person-centered and variable-centered approaches to longitudinal data. Merrill-Palmer Quarterly, 52(3), 377–389. doi: 10.1353/mpq.2006.0029 Leerkes, E., & Crockenberg, S. (2003). The impact of maternal characteristics and sensitivity on the concordance between maternal reports and laboratory observations of infant negative emotionality. Infancy, 4, 517–539.

Lemery, K., Essex, M., & Smider, N. (2002). Revealing the relation between temperament and behavior problem symptoms by eliminating measurement confounding: Expert ratings and factor analyses. Child Development, 73(3), 867–882. doi: 10.1111/1467–8624.00444 Lemery, K., Goldsmith, H., Klinnert, M., & Mrazek, D. (1999). Developmental models of infant and childhood temperament. Developmental Psychology, 35(1), 189–204. Lemery-Chalfant, K., Doelger, L., & Goldsmith, H. (2008). Genetic relations between effortful and attentional control and symptoms of psychopathology in middle childhood. Infant and Child Development, 17, 365–385. doi: 10.1002/icd.581 Lengua, L. (2006). Growth in temperament and parenting as predictors of adjustment during children's transition to adolescence. Development Psychology, 42, 819–832. doi: 10.1037/0012–1649.42.5.819 Lengua, L. (2008). Anxiousness, frustration, and effortful control as moderators of the relation between parenting and adjustment in middle childhood. Social Development, 17, 554–577. doi: 10.1111/j.1467–9507.2007.00438.x Lengua, L., Honorado, E., & Bush, N. (2007). Contextual risk and parenting as predictors of effortful control and social competence in preschool children. Journal of Applied Developmental Psychology, 28, 40–55. doi: 10.1016/j.appdev.2006.10.001 Lengua, L., & Kovacs, E. (2005). Bidirectional associations between temperament and parenting and the prediction of adjustment problems in middle childhood. Journal of Applied Developmental Psychology, 26, 21–38. doi: 10.1016/j.appdev.2004.10.001 Lengua, L., & Wachs, T. (2012). Temperament and risk: Resilient and vulnerable responses to adversity. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 519–540). New York, NY: Guilford Press. Lengua, L., West, S., & Sandler, I. (1998). Temperament as a predictor of symptomatology in children: Addressing contamination of measures. Child Development, 69, 164–181. Lengua, L., Wolchik, S., Sandler, I., & West, S. (2000). The additive and interactive effects of parenting and temperament in predicting problems of children of divorce. Journal of Clinical Child and Adolescent Psychology, 29, 232–244. doi: 10.1207/S15374424jccp2902_9 Leve, L., Kim, H., & Pears, K. (2005). Childhood temperament and family environment as predictors of internalizing and externalizing trajectories from ages 5 to 17. Journal of Abnormal Child Psychology, 33(5), 505–520. doi: 10.1007/s10802–005–6734–7 Lewis, M., Wolan-Sullivan, M., & Ramsay, D. S. (1992). Individual differences in anger and sad expressions during extinction: Antecedents and consequences. Infant Behavior and Development, 15, 443–452. Loeber, R., & Burke, J. (2011). Developmental pathways in juvenile externalizing and

internalizing problems. Journal of Research on Adolescence, 21(1), 34–46. doi: 10.1111/j.1532–7795.2010.00713.x Loken, E. (2004). Using latent class analysis to model temperament types. Multivariate Behavioral Research, 39(4), 625–652. Lonigan, C., Vasey, M., Phillips, B., & Hazen, R. (2004). Temperament, anxiety, and the processing of threat-relevant stimuli. Journal of Clinical Child and Adolescent Psychology, 33, 8–20. Lopez-Duran, N., Kovacs, M., & George, C. (2009). Hypothalamic-pituitary-adrenal axis dysregulation in depressed children and adolescents: A meta-analysis. Psychoneuroendocrinology, 34(9), 1272–1283. doi: 10.1016/j.psyneuen.2009.03.016 Luthar, S. S. (2006). Resilience in development: A synthesis of research across five decades. In D. C. Cohen & D. Cicchetti (Ed.), Developmental psychopathology, Vol. 3: Risk, disorder, and adaptation (2nd ed.) (pp. 739–795). Hoboken, NJ: Wiley. Magnusson, D. (1998). The logic and implications of a person-oriented approach. In R. B. Cairns, L. R. Bergman, & J. Kagan (Eds.), Methods and models for studying the individual (pp. 33–62). Thousand Oaks, CA: Sage. Magnusson, D., & Bergman, L. R. (1990). A pattern approach to the study of pthways from childhood to adulthood. In L. Robins & M. Rutter (Eds.), Straight and devious pathways from childhood to adulthood (pp. 101–115). Cambridge, UK: Cambridge University Press. Majdandzić, M., & van den Boom, D. C. (2007). Multimethod longitudinal assessment of temperament in early childhood. Journal of Personality, 75(1), 121–168. doi: 10.1111/j.1467–6494.2006.00435.x Mangelsdorf, S. C., Schoppe, S. J., & Buur, H. (2000). The meaning of parental reports: A contextual approach to the study of temperament and behavior problems in childhood. In V. J. Molfese, D. L. Molfese, & R. R. McCrae (Eds.), Temperament and personality development across the life span (pp. 121–140). Mahwah, NJ: Erlbaum. Marakovitz, S., Wagmiller, R., Mian, N., Briggs-Gowan, M., & Carter, A. (2011). Lost toy? Monsters under the bed? Contributions of temperament and family factors to early internalizing problems in boys and girls. Journal of Clinical Child and Adolescent Psychology, 40, 233– 244. doi: 10.1080/15374416.2011.546036 Marcovitch, S., Leigh, J., Calkins, S., Leerks, E., O'Brien, M., & Blankson, A. (2010). Moderate vagal withdrawal in 3.5-year-old children is associated with optimal performance on executive function tasks. Developmental Psychobiology, 52(6), 603–608. doi: 10.1002/dev.20462 Marsh, A., Finger, E., Mitchell, D., Reid, M., Sims, C., Kosson, D.,…Blair, R. (2008). Reduced amygdala responses to fearful expressions in children and adolescents with callous-

unemotional traits and disruptive behavior disorders. American Journal of Psychiatry, 165, 712–720. Martel, M., Nigg, J., Wong, M., Fitzgerald, H., Jester, J., Puttler, L., & Zucker, R. (2007). Childhood and adolescent resiliency, regulation, and executive functioning in relation to adolescent problems and competence in a high-risk sample. Development and Psychopathology, 19, 541–563. doi: 10.1017/S0954579407070265 Matheny, A. (1983). A longitudinal twin study of stability of components from Bayley's Infant Behavior Record. Child Development, 54, 356–360. Matthews, K., Salomon, K., Kenyon, K., & Allen, M. (2002). Stability of children's and adolescents' hemodynamic responses to psychological challenge: A three-year longitudinal study of a multiethnic cohort of boys and girls. Psychophysiology, 39(6), 826–834. doi: 10.1111/1469–8986.3960826 McAdams, D. P., & Pals, J. L. (2006). A new Big Five: Fundamental principles for an integrative science of personality. American Psychologist, 61(3), 204–217. doi: 10.1037/0003–066X.61.3.204 McClowry, S., Giangrande, S., Tommasini, N., Clinton, W., Foreman, N., Lynch, K., & Ferketich, S. (1994). The effects of child temperament, maternal characteristics, and family circumstances on the maladjustment of school-age children. Research in Nursing & Health, 17, 25–35. doi: 10.1002/nur.4770170105 McClowry, S., Snow, D., Tamis-Lemonda, C., & Rodrigues, E. (2010). Testing the efficacy of INSIGHTS on student disruptive behavior, classroom management, and students competence in inner city primary grades. School Mental Health, 2, 23–35. McCrae, R. R., Costa Jr, P. T., Ostendorf, F., Angleitner, A., Hřebíčková, M., Avia, M. D.,… Smith, P. B. (2000). Nature over nurture: Temperament, personality, and life span development. Journal of Personality and Social Psychology, 78(1), 173–186. doi: 10.1037/0022–3514.78.1.173 McCrory, E., DeBrito, S., & Viding, E. (2010). Research review: The neurobiology and genetics of maltreatment and adversity. Journal of Child Psychology and Psychiatry, 51, 1079–1095. McDermott, J., Perez-Edgar, K., Henderson, H., Chronis-Tuscano, A., Pine, D., & Fox, N. (2009). A history of childhood behavioral inhibition and enhanced response monitoring in adolescence are linked to clinical anxiety. Biological Psychiatry, 65(5), 445–448. doi: 10.1016/j.biopsych.2008.10.043 McDevitt, S., & Carey, W. (1978). The measurement of temperament in 3–7 year old children. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 19(3), 245–253. McGrath, J., & O'Brien, W. (2001). Pediatric impedance cardiography: Temporal stability and

intertask consistency. Psychophysiology, 38, 479–484. McGrath, J. M., Records, K., & Rice, M. (2008). Maternal depression and infant temperament characteristics. Infant Behavior & Development, 31(1), 71–80. doi: 10.1016/j.infbeh.2007.07.001 Mervielde, I., & De Pauw, S. (2012). Models of child temperament. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 21–40). New York, NY: Guilford Press. Mesman, J., & Koot, H. (2001). Early preschool predictors of preadolescent internalizing and externalizing DSM-IV diagnoses. American Academy of Child & Adolescent Psychiatry, 40, 1029–1036. doi: 10.1097/00004583–200109000–00011 Mesman, J., Stoel, R., Bakermans-Kranenburg, M., van IJzendoorn, M., Juffer, F., Koot, H., & Alink, L. (2009). Predicting growth curves of early childhood externalizing problems: Differential susceptibility of children with difficult temperament. Journal of Abnormal Child Psychology, 37, 625–636. doi: 10.1007/s10802–009–9298–0 Messinger, D. S., Mahoor, M. H., Chow, S.-M., & Cohn, J. F. (2009). Automated measurement of facial expression in infant-mother interaction: A pilot study. Infancy, 14(3), 285–305. doi: 10.1080/15250000902839963 Messinger, D. S., Mattson, W. I., Mahoor, M. H., & Cohn, J. F. (2012). The eyes have it: Making positive expressions more positive and negative expressions more negative. Emotion, 12(3), 430–436. doi: 10.1037/a0026498 Miskovic, V., & Schmidt, L. (2012). New directions in the study of individual differences in temperament: A brain-body approach to understanding fearful and fearless children. Monographs of the Society for Research in Child Development, 77, 28–38. Moffit, T., Caspi, A., & Rutter, M. (2006). Measured gene-environment interactions in psychopathology: Concepts, research strategies, and implications for research, intervention, and public understanding of genetics. Psychological Science, 1, 5–27. Morris, A., Silk, J., Steinberg, L., Sessa, F., Avenevoli, S., & Essex, M. (2002). Temperamental vulnerability and negative parenting as interacting of child adjustment. Journal of Marriage and Family, 64, 461–471. doi: 10.1111/j.1741-3737.2002.00461.x Morrissey, T. (2009). Multiple child-care arrangements and young children's behavioral outcomes. Child Development, 80, 59–76. doi: 10.1111/j.1467–8624.2008.01246.x Mullineaux, P., Deater-Deckard, K., Petrill, S., Thompson, L., & DeThorne, L. (2009). Temperament in middle childhood: A behavioral genetic analysis of fathers' and mothers' reports. Journal of Research in Personality, 43, 737–746. Muris, P., Meesters, C., & Blijlevens, P. (2007). Self-reported reactive and regulative temperament in early adolescence: Relations to internalizing and externalizing problem

behavior and “Big Three” personality factors. Journal of Adolescence, 30, 1035–1049. doi: 10.1016/j.adolescence.2007.03.003 Muris, P., van der Pennen, E., Sigmond, R., & Mayer, B. (2008). Symptoms of anxiety, depression, and aggression in non-clinical children: Relationships with self-report and performance-based measures of attention and effortful control. Child Psychiatry and Human Development, 39, 455–467. doi: 10.1007/s10578–008–0101–1 Murray, K. T., & Kochanska, G. (2002). Effortful control: Factor structure and relation to externalizing and internalizing behaviors. Journal of Abnormal Child Psychology, 30(5), 503– 514. Nachmias, M., Gunnar, M., Mangelsdorf, S., Parritz, R., & Buss, K. (1996). Behavioral inhibition and stress reactivity: the moderating role of attachment security. Child Development, 67(2), 508–522. Nigg, J. (2001). Is ADHD a disinhibitory disorder? Psychological Bulletin, 127, 571–598. doi: 10.1037/0033–2909.127.5.571 Nigg, J. (2006). Temperament and developmental psychopathology. Journal of Child Psychology and Psychiatry, 47, 395–422. doi: 10.1111/j.1469–7610.2006.01612.x Odgers, C., Moffitt, T., Broadbent, J., Dickson, N., Hancox, R., Harrington, H.,…Caspi, A. (2008). Female and male antisocial trajectories: From childhood origins to adult outcomes. Development and Psychopathology, 20(2), 673–716. doi: 10.1017/S0954579408000333 Oldehinkel, A., Hartman, C., De Winter, A., Veenstra, R., & Ormel, J. (2004). Temperament profiles associated with internalizing and externalizing problems in preadolescence. Development and Psychopathology, 16, 421–440. doi: 10.1017/S0954579404044591 Oldehinkel, A., Hartman, C., Ferdinand, R., Verhulst, F., & Ormel, J. (2007). Effortful control as modifier of the association between negative emotionality and adolescents' mental health problems. Development and Psychopathology, 19, 523–539. doi: 10.1017/S0954579407070253 Olino, T., Klein, D., Dyson, M., Rose, S., & Durbin, E. (2010). Temperamental emotionality in preschool-aged children and depressive disorders in parents: Associations in a large community sample. Journal of Abnormal Child Psychology, 119, 468–478. doi: 10.1037/a0020112 Ollendick, T., & King, N. (1994). Diagnosis, assessment, and treatment of internalizing problems in children: The role of longitudinal data. Journal of Consulting and Clinical Psychology, 62(5), 918–927. doi: 10.1037/0022–006X.62.5.918 Olson, S., Bates, J., Sandy, J., & Lanthier, R. (2000). Early developmental precursors of externalizing behavior in middle childhood and adolescence. Journal of Abnormal Child Psychology, 28, 119–133. doi: 10.1023/A:1005166629744

Olson, S., Sameroff, A., Kerr, D., Lopez, N., & Wellman, H. (2005). Developmental foundations of externalizing problems in young children: The role of effortful control. Development and Psychopathology, 17, 25–45. doi: 10.1017/S0954579405050029 Owens, E., & Shaw, D. (2003). Predicting growth curves of externalizing behavior across the preschool years. Journal of Abnormal Child Psychology, 31, 575–590. doi: 10.1023/A:1026254005632 Park, S., Belsky, J., Putnam, S., & Crnic, K. (1997). Infant emotionality, parenting, and 3-year inhibition: Exploring stability and lawful discontinuity in a male sample. Developmental Psychology, 33, 218–227. Patterson, G. (1976). The aggressive child: Victim and architect of a coercive system. In E. Mash, L. Hamerlynck, & L. Handy (Eds.), Behavior modification and families (pp. 267–316). New York, NY: Brunner/Mazel. Patterson, G. (1982). Coercive family process. Eugene, OR: Castalia. Paulussen-Hoogeboom, M., Stams, G., Hermanns, J., & Peetsma, T. (2007). Child negative emotionality and parenting from infancy to preschool: A meta-analytic review. Developmental Psychology, 43, 438–453. doi: 10.1037/0012–1649.43.2.438 Perez-Edgar, K., Bar-Haim, Y., McDermott, J., Chronis-Tuscano, A., Pine, D., & Fox, N. (2010). Attention biases to threat and behavioral inhibition in early childhood shape adolescent social withdrawal. Emotion, 10, 349–357. doi: 10.1037/a0018486 Perez-Edgar, K., Bar-Haim, Y., McDermott, J., Gorodetsky, E., Hodgkinson, C., Goldman, D., & Fox, N. (2010). Variations in the serotonin-transporter gene are associated with attention bias patterns to positive and negative emotion faces. Biological Psychiatry, 83, 269–271. doi: 10.1016/j.biopsycho.2009.08.009 Perez-Edgar, K., & Fox, N. (2005a). A behavioral and electrophysiological study of children's selective attention under neutral and affective conditions. Journal of Cognition and Development, 6, 89–118. doi: 10.1207/s15327647jcd0601_6 Perez-Edgar, K., & Fox, N. (2005b). Temperament and anxiety disorders. Child and Adolescent Psychiatric Clinics of North America, 14(4), 681–706. doi: 10.1016/j.chc.2005.05.008 Perez-Edgar, K., Robertson-Nay, R., Hardin, M., Poeth, K., Guyer, A., Nelson, E.,…Ernst, M. (2007). Attention alters neural repsonses to evocative faces in behaviorally inhibited adolescents. NeuroImage, 35, 1538–1546. Perez-Edgar, K., Schmidt, L., Henderson, H., Schulkin, J., & Fox, N. (2008). Salivary cortisol levels and infant temperament shape developmental trajectories in boys at risk for behavioral maladjustment. Psychoneuroendocrinology, 33(7), 916–925. doi: 10.1016/j.psyneuen.2008.03.018

Pfeifer, M., Goldsmith, H., Davidson, R., & Rickman, M. (2002). Continuity and change in inhibited and uninhibited children. Child Development, 73(5), 1474–1485. Pine, D., Cohen, P., Gurley, D., Brook, J., & Ma, Y. (1998). The risk for early-adulthood anxiety and depressive disorders in adolescents with anxiety and depressive disorders. Archives of General Psychiatry, 55(1), 56–64. doi: 10.1001/archpsyc.55.1.56 Plomin, R. (1982). The difficult concept of temperament: A response to Thomas, Chess, and Korn. Merrill-Palmer Quarterly, 28, 25–33. Polak-Toste, C., & Gunnar, M. (2006). Temperamental exuberance: Correlates and consequences. In P. Marshall & N. Fox (Eds.), The development of social engagement: Neurobiological perspectives (19–45). Oxford, UK: Oxford University Press. Porges, S. (1992). Vagal tone: A physiologic marker of stress vulnerability. Pediatrics, 90(3), 498–504. Porges, S. (2001). The polyvagal theory: Phylogenetic substrates of a social nervous system. International Journal of Psychophysiology, 42(2), 123–146. doi: 10.1016/S0167– 8760(01)00162–3 Porges, S. (2003). The polyvagal theory: Phylogenetic contributions to social behavior. Physiology & Behavior, 79(3), 503–513. doi: 10.1016/S0031–9384(03)00156–2 Porges, S. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–143. doi: 10.1016/j.biopsycho.2006.06.009 Porges, S., Doussard-Roosevelt, J., & Maiti, A. (1994). Vagal tone and the physiological regulation of emotion. Monographs of the Society for Research in Child Development, 59(2– 3), 167–186. doi: 10.1111/j.1540–5834.1994.tb01283.x Porges, S., Doussard-Roosevelt, J., Portales, A., & Greenspan, S. (1996). Infant regulation of the vagal “brake” predicts child behavior problems: A psychobiological model of social behavior. Developmental Psychobiology, 29(8), 697–712. doi: 10.1002/(SICI)1098– 2302(199612)29:83.0.CO;2-O Posner, M., & Rothbart, M. (2000). Developing mechanisms of self-regulation. Development and Psychopathology, 12(3), 427–441. Posner, M., & Rothbart, M. (2007). Research on attention networks as a model for the integration of psychological science. Annual Review of Psychology, 58, 1–23. Propper, C., & Moore, G. (2006). The influence of parenting on infant emotionality: A multilevel psychobiological perspective. Developmental Review, 26(4), 427–460. doi: 10.1016/j.dr.2006.06.003 Putnam, S. (2012). Positive emotionality. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 105–123). New York, NY: Guilford Press.

Putnam, S., Ellis, L., & Rothbart, M. (2001). The structure of temperament from infancy through adolescence. In A. Eliasz & A. Angleitner (Eds.), Advances in research on temperament (pp. 165–182). Lengerich, Germany: Pabst Science. Putnam, S., Gartstein, M., & Rothbart, M. (2006). Measurement of fine-grained aspects of toddler temperament: The Early Childhood Behavior Questionnaire. Infant Behavior and Development, 29, 386–401. Putnam, S., Rothbart, M., & Gartstein, M. (2008). Homotypic and heterotypic continuity of fine-grained temperament during infancy, toddlerhood, and early childhood. Infant and Child Development, 17, 387–405. Putnam, S., Sanson, A., & Rothbart, M. (2002). Child temperament and parenting. In M. Bornstein (Ed.), Handbook of parenting: Vol. 1: Children and parenting (2nd ed., pp. 255– 277). Mahwah, NJ: Erlbaum. Putnam, S., & Stifter, C. (2002). Development of approach and inhibition in the first year: Parallel findings from motor behavior, temperament ratings and directional cardiac response. Developmental Science, 5(4), 441–451. doi: 10.1111/1467–7687.00239 Putnam, S., & Stifter, C. (2005). Behavioral approach–inhibition in toddlers: Prediction from infancy, positive and negative affective components, and relations with behavior problems. Child Development, 76(1), 212–226. doi: 10.1111/j.1467–8624.2005.00840.x Quigley, K., & Stifter, C. (2006). A comparative validation of sympathetic reactivity in children and adults. Psychophysiology, 43, 357–365. Raine, A. (1993). The psychopathology of crime: Criminal behavior as a clinical disorder. San Diego, CA: Academic Press. Rapee, R., Kennedy, S., Ingram, M., Edwards, S., & Sweeney, L. (2010). Altering the trajectory of anxiety in at-risk young children. American Journal of Psychiatry, 167, 1518– 1525. Rhoades, K., Leve, L., Harold, G., Neiderhiser, J., Shaw, D., & Reiss, D. (2011). Longitudinal pathways from marital hostility to child anger during toddlerhood: Genetic susceptibility and indirect effects via harsh parenting. Journal of Family Psychology, 25, 282–291. doi: 10.1037/a0022886 Rimm-Kaufman, S., & Kagan, J. (2005). Infant predictors of kindergarten behavior: The contribution of inhibited and uninhibited temperament types. Behavioral Disorders, 30, 331– 347. Roberts, B. W., & DelVecchio, W. F. (2000). The rank-order consistency of personality traits from childhood to old age: A quantitative review of longitudinal studies. Psychological Bulletin, 126(1), 3–25. doi: 10.1037/0033–2909.126.1.3

Robins, R., John, O., & Caspi, A. (1998). The typological approach to studying personality. In R. B. Cairns, L. R. Bergman, & J. Kagan (Eds.), Methods and models for studying the individual (pp. 135–160). Thousand Oaks, CA: Sage. Root, A., & Stifter, C. (2010). Temperament and maternal emotion socialization beliefs as predictors of early childhood social behavior in the laboratory and classroom. Parenting Science and Practice, 10, 241–257. doi: 10.1080/15295192.2010.492035 Rosen, J., & Schulkin, J. (1998). From normal fear to pathological anxiety. Psychological Review, 105, 325–350. Rothbart, M. (1981). Measurement of temperament in infancy. Child Development, 52, 569– 578. Rothbart, M. (1982). The concept of difficult temperament: A critical analysis of Thomas, Chess, and Korn. Merrill-Palmer Quarterly, 28, 35–40. Rothbart, M. (1988). Temperament and the development of inhibited approach. Child Development, 59, 1241–1250. Rothbart, M. (1989). Temperament and development. In G. Kohnstamm, J. Bates, & M. Rothbart (Eds.), Temperament in childhood (pp. 187–247). New York, NY: Wiley. Rothbart, M. (2004). Commentary: Differentiated measures of temperament and multiple pathways to childhood disorders. Journal of Clinical Child and Adolescent Psychology, 33, 82–87. doi: 10.1207/S15374424JCCP3301_8 Rothbart, M. (2011). Becoming who we are: Temperament and personality in development. New York, NY: Guilford Press. Rothbart, M. (2012). Advances in temperament: History, concepts, and measures. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 3–20). New York, NY: Guilford Press. Rothbart, M., & Ahadi, S. (1994). Temperament and the development of personality. Journal of Abnormal Psychology, 103, 55–66. doi: 10.1037/0021–843X.103.1.55 Rothbart, M., Ahadi, S., & Evans, D. (2000). Temperament and personality: Origins and outcomes. Journal of Personality and Social Psychology, 78(1), 122–135. Rothbart, M., Ahadi, S., & Hershey, K. (1994). Temperament and social behavior in childhood. Merrill-Palmer Quarterly, 40(1), 21–39. Rothbart, M., Ahadi, S., Hershey, K., & Fisher, P. (2001). Investigations of temperament at three to seven years: The Children's Behavior Questionnaire. Child Development, 72(5), 1394–1408. Rothbart, M., & Bates, J. ( 2006). Temperament. In R. L. W. Damon, & N. Eisenberg (Ed.),

Handbook of child psychology, Sixth edition: Social, emotional, and personality development (Vol. 3, pp. 99–166). Hoboken, NJ: Wiley. Rothbart, M., Chew, K. H., & Gartstein, M. (2001). Assessment of temperament in early development. In L. T. Singer & P. S. Zeskind (Eds.), Biobehavioral assessment of the infant (pp. 190–208). New York, NY: Guilford Press. Rothbart, M., & Derryberry, D. (1981). Development of individual differences in temperament. In M. E. Lamb & A. L. Brown (Eds.), Advances in developmental psychology (Vol. 1, pp. 37– 86). Hillsdale, NJ: Erlbaum. Rothbart, M., Derryberry, D., & Hershey, K. (2000). Stability of temperament in childhood: Laboratory infant assessment to parent report at seven years. In V. J. Molfese & D. L. Molfese (Eds.), Temperament and personality development across the life span (85–119). Mahwah, NJ: Erlbaum. Rothbart, M., Derryberry, D., & Posner, M. (1994). A psychobiological approach to the development of temperament. In J. Bates & T. Wachs (Eds.), Temperament: Individual differences at the interface of biology and behavior (pp. 83–116). Washington, DC: American Psychological Association. Rothbart, M., & Goldsmith, H. (1985). Three approaches to the study of infant temperament. Developmental Review, 5, 237–260. Rothbart, M., & Posner, M. (2006). Temperament, attention, and developmental psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Developmental neuroscience (2nd ed., Vol. 2, pp. 465–501). Hoboken, NJ: Wiley. Rothbart, M., Posner, M., & Hershey, K. (1995). Temperament, attention, and developmental psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Handbook of developmental psychopathology (Vol. 1, pp. 315–340). New York, NY: Wiley. Rothbart, M., & Rueda, M. (2005). The development of effortful control. In U. Mayr, E. Awh, & S. W. Keele (Eds.), Developing individuality in the human brain: A tribute to Michael I. Posner (pp. 167–188). Washington, DC: American Psychological Association. Rowe, D., & Kandel, D. (1997). In the eye of the beholder? Parental ratings of externalizing and internalizing symptoms. Journal of Abnormal Child Psychology, 25(4), 265–275. Rowe, D., & Plomin, R. (1977). Temperament in early childhood. Journal of Personality Assessment, 41(2), 150–156. Rubin, K., Bukowski, W., & Parker, J. (2006). Peer interactions, relationships, and groups. In N. Eisenberg, W. Damon, & R. Lerner (Eds.), Handbook of child psychology: Vol. 3, Social, emotional, and personality development (6th ed., pp. 571–645). Hoboken, NJ: Wiley. Rubin, K., Burgess, K., & Hastings, P. (2002). Stability and social-behavioral consequences of

toddlers' inhibited temperament and parenting behaviors. Child Development, 73, 483–495. Rubin, K., Cheah, C., & Fox, N. (2001). Emotion regulation, parenting and display of social reticence in preschoolers. Early Education and Development, 12, 97–115. doi: 10.1207/s15566935eed1201_6 Rubin, K., Coplan, R., Fox, N., & Calkins, S. (1995). Emotionality, emotion regulation, and preschoolers' social adaptation. Development and Psychopathology, 7, 49–62. Rubin, K., Hastings, P., Stewart, S., Henderson, H., & Chen, X. (1997). The consistency and concomitants of inhibition: Some of the children, all of the time. Child Development, 68, 467– 483. doi: 10.2307/1131672 Rudea, M. R. (2012). Effortful control. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 145–167). New York, NY: Guilford Press. Rutter, M. (1986). Child psychiatry: The interface between clinical and developmental research. Psychological Medicine, 16, 151–169. doi: 10.1017/S0033291700002592 Rydell, A. M., Berlin, L., & Bohlin, G. (2003). Emotionality, emotion regulation, and adaptation among 5- to 8-year old children. Emotion, 3(1), 30–47. Saarni, C., Mumme, D., & Campos, J. (1998). Emotional development: Action, communication, and understanding. In N. Eisenberg (Ed.), Handbook of child psychology, Vol 3. Social, emotional, and personality development (5th ed., pp. 237–309). New York, NY: Wiley. Sameroff, A. (2009). The transactional model of development: How children and contexts shape each other. Washington, DC: American Psychological Association. Sameroff, A., Seifer, R., & Elias, P. (1982). Sociocultural variability in infant temperament ratings. Child Development, 53, 164–173. Sanson, A., Prior, M., & Kyrios, M. (1990). Contamination of measures in temperament research. Merrill-Palmer Quarterly, 36(2), 179–192. Saudino, K. (2009). Do different measures tap the same genetic influences? A multi-method study of activity level in young twins. Developmental Science, 12(4), 626–633. doi: 10.1111/j.1467–7687.2008.00801.x Saudino, K., & Cherny, S. (2001). Sources of continuity and change in observed temperament. In R. N. Emde & J. K. Hewitt (Eds.), Infancy to early childhood: Genetic and environmental influences on developmental change (pp. 89–110). New York, NY: Oxford University Press. Saudino, K., & Eaton, W. (1991). Infant temperament and genetics: An objective twin study of motor activity level. Child Development, 62(5), 1167–1174. Saudino, K., & Wang, M. (2012). Quantitative and molecular genetic studies of temperament.

In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 315–346). New York, NY: Guilford Press. Saudino, K., Wertz, A., Gagne, J., & Chawla, S. (2004). Night and day: Are siblings as different in temperament as parents say they are? Journal of Personality and Social Psychology, 87(5), 698–706. doi: 10.1037/0022–3514.87.5.698 Schmidt, L. A., & Fox, N. (1994). Patterns of cortical electrophysiology and autonomic activity in adults' shyness and sociability. Biological Psychology, 38(2–3), 183–198. doi: 10.1016/0301–0511(94)90038–8 Schmidt, L. A., Fox, N. A., Schulkin, J., & Gold, P. W. (1999). Behavioral and psychophysiological correlates of self-presentation in temperamentally shy children. Developmental Psychobiology, 35, 119–135. Schmitz, S., Fulker, D., Plomin, R., Zahn-Waxler, C., Emde, R., & DeFries, J. (1999). Temperament and problem behavior during early childhood. International Journal of Behavioral Development, 23, 333–355. Schwartz, C., Snidman, N., & Kagan, J. (1996). Early childhood temperament as a determinant of externalizing behavior in adolescence. Development and Psychopathology, 8, 527–537. Schwartz, C., Snidman, N., & Kagan, J. (1999). Adolescent social anxiety as an outcome of inhibited temperament in childhood. Journal of the American Academy of Child and Adolescent Psychiatry, 38, 1008–1015. doi: 10.1097/00004583–199908000–00017 Schwartz, C., Wright, C., Shin, L., Kagan, J., & Rauch, S. (2003). Inhibited and unihibited infants “grown up”: Adult amygdalar response to novelty. Science, 300, 1952–1953. Seifer, R., Sameroff, A., Barrett, L., & Krafchuk, E. (1994). Infant temperament measured by multiple observations and mother report. Child Development, 65, 1478–1490. Seifer, R., Sameroff, A., Dickstein, S., Schiller, M., & Hayden, L. (2004). Your own children are special: Clues to the sources of reporting bias in temperament assessments. Infant Behavior and Development, 27, 323–341. Shankman, S., Klein, D., Torpey, D., Olino, T., Dyson, M., Kim, J., & Tenke, C. (2011). Do positive and negative temperament traits interact in predicting risk for depression? A resting EEG study of 329 preschoolers. Development and Psychopathology, 23, 551–562. doi: 10.1017/S0954579411000022 Shankman, S., Tenke, C., Bruder, G., Durbin, E., Hayden, E., & Klein, D. (2005). Low positive emotionality in young children: Association with EEG asymmetry. Development and Psychopathology, 17, 85–98. doi: 10.1017/S0954579405050054 Sheese, B., Voelker, P., Rothbart, M., & Posner, M. (2007). Parenting quality interacts with genetic variation in dopamine receptor D4 to influence temperament in early childhood.

Development and Psychopathology, 19(4), 1039–1046. doi: 10.1017/S0954579407000521 Shiner, R., Buss, K., McClowry, S., Putnam, S., Saudino, K., & Zentner, M. (2012). What is temperament now? Assessing progress in temperament research on the twenty-fifth anniversary of Goldsmith et al. (1987). Child Development Perspectives, 6(4), 436–444. doi: 10.1111/j.1750–8606.2012.00254.x Shiner, R., & Caspi, A. (2012). Temperament and the development of personality traits, adaptations, and narratives. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 497–516). New York, NY: Guilford Press. Shiner, R., & Masten, A. (2008). Personality in childhood: A bridge from early temperament to adult outcomes. International Journal of Developmental Science, 2(1), 158–175. doi: 10.3233/DEV-2008–21210 Smith, A., Rhee, S., Corley, R., Friedman, N., Hewitt, J., & Robinson, J. (2012). The magnitude of genetic and environmental influences on parental and observational measures of behavioral inhibition and shyness in toddlerhood. Behavioral Genetics, 42(5), 764–777. doi: 10.1007/s10519–012–9551–0 Soto, C., John, O., Gosling, S., & Potter, J. (2011). Age differences in personality traits from 10 to 65: Big Five domains and facets in a large cross-sectional sample. Journal of Personality and Social Psychology, 100(2), 330–348. doi: 10.1037/a0021717 Source, J., & Emde, R. (1982). The meaning of infant emotional expressions: Regularities in caregiving responses in normal and Down's syndrome infants. Journal of Child Psychology and Psychiatry, 23, 145–158. Spinrad, T., Eisenberg, N., Gaertner, B., Popp, T., Smith, C., Kupfer, A., & Hofer, C. (2007). Relations of maternal socialization and toddlers' effortful control to children's adjustment and social competence. Developmental Psychology, 43, 1170–1186. doi: 10.1037/0012– 1649.43.5.1170 Spinrad, T., Eisenberg, N., Granger, D., Eggum, N., Sallquist, J., Haugen, R.,…Hofer, C. (2009). Individual differences in preschoolers' salivary cortisol and alpha-amylase reactivity: relations to temperament and maladjustment. Hormones and Behavior, 56(1), 133–139. doi: 10.1016/j.yhbeh.2009.03.020 Sroufe, A. (1990). Considering normal and abnormal together: The essence of developmental psychopathology. Development and Psychopathology, 2, 335–347. Sroufe, A. (2009). The concept of development in developmental psychopathology. Child Development Perspectives, 3(3), 178–183. Sroufe, A., & Rutter, M. (1984). The domain of developmental psychopathology. Child Development, 55, 17–29.

Stadler, C., Sterzer, P., Schmeck, K., Krebs, A., Kleinschmidt, A., & Poustka, F. (2007). Reduced anterior cingulate activation in aggressive children and adolescents during affective stimulation: Association with temperament traits. Journal of Psychiatric Research, 41, 410– 417. Stifter, C., & Corey, J. (2001). Vagal regulation and observed social behavior in infancy. Social Development, 10, 189–201. Stifter, C., Dollar, J., & Cipriano, E. (2011). Temperament and emotion regulation: the role of autonomic nervous system reactivity. Developmental Psychobiology, 53(3), 266–279. doi: 10.1002/dev.20519 Stifter, C., & Fox, N. (1990). Infant reactivity: Physiological correlates of newborn and five month temperament. Developmental Psychology, 26, 582–588. Stifter, C., Fox, N., & Porges, S. (1989). Facial expressivity and vagal tone in 5- and 10month-old infants. Infant Behavior and Development, 12(2), 127–137. doi: 10.1016/0163– 6383(89)90001–5 Stifter, C., & Jain, A. (1996). Psychophysiological correlates of infant temperament: Stability of behavior and autonomic patterning from 5 to 18 months. Developmental Psychobiology, 29(4), 379–391. doi: 10.1002/(SICI)1098–2302(199605)29:43.0.CO;2-N Stifter, C., Putnam, S., & Jahromi, L. (2008). Exuberant and inhibited toddlers: Stability of temperament and prediction to behavior problems. Development and Psychopathology, 20, 401–421. Stifter, C., Willoughby, M., & Towe-Goodman, N. (2008). Agree or agree to disagree? Assessing the convergence between parents and observers on infant temperament. Infant and Child Development, 17, 407–426. Stringaris, A., Maughan, B., & Goodman, R. (2010). What is a disruptive disorder? Temperamental antecedents of oppositional defiant disorder: Findings from the Avon Longitudinal Study. Journal of the American Academy of Child and Adolescent Psychiatry, 49, 474–483. doi: 10.1097/00004583–201005000–00008 Suess, P., Porges, S., & Plude, D. (1994). Cardiac vagal tone and sustained attention in schoolage children. Psychophysiology, 31(1), 17–22. doi: 10.1111/j.1469–8986.1994.tb01020.x Tackett, J. (2006). Evaluating models of the personality–psychopathology relationship in children and adolescents. Clinical Psychology Review, 26(5), 584–599. doi: 10.1016/j.cpr.2006.04.003 Tackett, J., Martel, M., & Kushner, S. (2012). Temperament, externalizing disorders, and attention-deficit/hyperactivity disorder. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 562–580). New York, NY: Guilford Press.

Tarullo, A., Mliner, S., & Gunnar, M. (2011). Inhibition and exuberance in preschool classrooms: associations with peer social experiences and changes in cortisol across the preschool year. Developmental Psychology, 47(5), 1374–1388. doi: 10.1037/a0024093 Thomas, A., & Chess, S. (1977). Temperament and development. New York, NY: New York University Press. Thomas, A., Chess, S., & Birch, H. (1968). Temperament and behavior disorders in children. New York, NY: New York University Press. Thomas, A., Chess, S., Birch, H., Hertzig, M., & Korn, S. (1963). Behavioral individuality in early childhood. New York, NY: New York University Press. Thorell, L., Bohlin, G., & Rydell, A. (2004). Two types of inhibitory control: Predictive relations to social functioning. International Journal of Behavioral Development, 28, 193– 203. doi: 10.1080/01650250344000389 Torpey, D., Hajcak, G., Kim, J., Kujawa, A., Dyson, M., Olino, T., & Klein, D. (2013). Errorrelated brain activity in young children: Associations with parental anxiety and child temperamental negative emotionality. Journal of Child Psychology and Psychiatry, 54(8), 854–862. doi: 10.1111/jcpp.12041 Towe-Goodman, N., Stifter, C., Coccia, M., & Cox, M. (2011). Interparental aggression, attention skills, and early childhood behavior problems. Development and Psychopathology, 23, 563–576. Valiente, C., Eisenberg, N., Smith, C., Reiser, M., Fabes, R., Losoya, S., & Murphy, B. (2003). The relations of effortful control and reactive control to children's externalizing problems: A longitudinal assessment. Journal of Personality, 71, 1171–1196. doi: 10.1111/1467– 6494.7106011 Valiente, C., Eisenberg, N., Spinrad, T., Reiser, M., Cumberland, A., Losoya, S., & Liew, J. (2006). Relations among mothers' expressivity, children's effortful control, and their problem behaviors: A four-year longitudinal study. Emotion, 6, 459–472. doi: 10.1037/1528– 3542.6.3.459 van der Meere, J. (2002). The role of attention. In S. Sandberg (Ed.), Hyperactivity and attention disorders of childhood (2nd ed., pp. 162–213). Cambridge, UK: Cambridge University Press. van Goozen, S., Fairchild, G., Snoek, H., & Harold, G. (2007). The evidence for a neurobiological model of childhood antisocial behavior. Psychological Bulletin, 133(1), 149– 182. doi: 10.1037/0033–2909.133.1.149 van Ijzendoorn, M., & Bakermans-Kranenburg, M. (2012a). Differential susceptibility experiments: Going beyond correlational evidence: Comment on beyond mental health, differential susceptibility articles. Developmental Psychology, 48(3), 769–774. doi:

10.1037/a0027536 van Ijzendoorn, M., & Bakermans-Kranenburg, M. (2012b). Integrating temperament and attachment: The differential susceptibility paradigm. In M. Zentner & R. Shiner (Eds.), Handbook of temperament (pp. 403–424). New York, NY: Guilford Press. Van Leeuwen, K., Mervielde, I., Braet, C., & Bosmans, G. (2004). Child personality and parental behavior as moderators of problem behavior: variable- and person-centered approaches. Developmental Psychology, 40, 1028–1046. doi: 10.1037/0012–1649.40.6.1028 van Zeijl, J., Mesman, J., Stolk, M., Alink, L., Van IJzensdoorn, M., Bakermans-Kranenburg, M.,…Koot, H. (2007). Differential susceptibility to discipline: The moderating effect of child temperament on the association between maternal discipline and early childhood externalizing problems. Journal of Family Psychology, 21, 626–636. Vaughn, B., Bradley, C., Joffe, L., Seifer, R., & Barglow, P. (1987). Maternal characteristics measured prenatally are predicitive of ratings of temperamental “difficulty” on the Carey Infant Temperament Questionnaire. Developmental Psychology, 23, 152–161. Vermunt, J., & Magidson, J. (2002). Latent class cluster analysis. In J. Hagenaars & A. McCuthcheon (Eds.), Applied latent class analysis (pp. 89–106). Cambridge, UK: Cambridge University Press. Voegtline, K., & Stifter, C. (2010). Late-preterm birth, maternal symptomatology, and infant negativity. Infant Behavior & Development, 33(4), 545–554. Wachs, T. D. (1991). Environmental considerations in studies with non-extreme groups. In R. Plomin & T. D. Wachs (Eds.), Conceptualization and measurement of organism— Environment interaction (pp. 44–67). Washington, DC: American Psychological Association. Watamura, S., Donzella, B., Alwin, J., & Gunnar, M. (2003). Morning-to-afternoon increases in cortisol concentrations for infants and toddlers at child care: age differences and behavioral correlates. Child Development, 74(4), 1006–1020. Welsh, M. C., Pennington, B. F., & Groisser, D. B. (1991). A normative-developmental study of executive function: A window on prefrontal function in children. Developmental Neuropsychology, 7(2), 131–149. doi: 10.1080/87565649109540483 White, J. (1999). Review personality, temperament and ADHD: A review of the literature. Personality and Individual Differences, 27(4), 589–598. doi: 10.1016/S0191– 8869(98)00273–6 White, L., McDermott, J., Degnan, K., Henderson, H., & Fox, N. (2011). Behavioral inhibition and anxiety: The moderating roles of inhibitory control and attention shifting. Journal of Abnormal Child Psychology, 39, 735–747. doi: 10.1007/s10802–011–9490-x Whittle, S., Allen, N., Lubman, D., & Yucel, M. (2006). The neurobiological basis of

temperament: Towards a better understanding of psychopathology. Neuroscience and Biobehavioral Reviews, 30, 511–525. Willcutt, E., Doyle, A., Nigg, J., Faraone, S., & Pennington, B. (2005). Validity of the executive function theory of attention-deficit/hyperactivity disorder: A meta-analytic review. Biological Psychiatry, 57, 1336–1346. doi: 10.1016/j.biopsych.2005.02.006 Wolff, J., & Ollendick, T. (2006). The comorbidity of conduct problems and depression in childhood and adolescence. Clinical Child and Family Psychology Review, 9(3–4), 201–220. doi: 10.1007/s10567–006–0011–3 Wood, A., Rijsdijk, F., Saudino, K., Asherson, P., & Kuntsi, J. (2008). High heritability for a composite index of children's activity level measures. Behavior Genetics, 38(3), 266–276. doi: 10.1007/s10519–008–9196–1 Wood, A., Saudino, K., Rogers, H., Asherson, P., & Kuntsi, J. (2007). Genetic influences on mechanically-assessed activity level in children. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 48(7), 695–702. Wood, J., McLeod, B., Sigman, M., Hwang, W., & Chu, B. (2003). Parenting and childhood anxiety: Theory, empirical findings, and future directions. Journal of Child Psychology and Psychiatry, 44, 134–151. doi: 10.1111/1469–7610.00106 Zentner, M., & Bates, J. (2008). Child temperament: An integrative review of concepts, research programs, and measures. International Journal of Developmental Science, 2(1), 7– 37. Zetner, M., & Shiner, R. (2012). Fifty years of progress in temperament research: A synthesis of major themes, findings, and challenges and a look forward. In M. Zentner & R. Shiner (Eds.), Handbook of temperament. New York, NY: Guilford Press. Zhou, Q., Chen, S. H., & Main, A. (2012). Commonalities and differences in the research on children's effortful control and executive function: A call for an integrated model of selfregulation. Child Development Perspectives, 6(2), 112–121. doi: 10.1111/j.1750– 8606.2011.00176.x

Chapter 12 Interparental Conflict and Child Adjustment Ernest N. Jouriles, Renee McDonald, and Chrystyna D. Kouros INTRODUCTION AND HISTORY INTERPARENTAL CONFLICT AND CHILD ADJUSTMENT PROBLEMS Topic of the Conflict Behavioral Tactics Emotions Resolution of Conflict Summary on Dimensions of Interparental Conflict A Note about Frequency Developmental Considerations Prevalence and Degree of Children's Exposure to Interparental Conflict THEORY EXPLAINING THE LINK BETWEEN INTERPARENTAL CONFLICT AND CHILD ADJUSTMENT Direct-Effects Models Indirect-Effects Models CHILD ATTRIBUTES AND FAMILY/COMMUNITY FACTORS CONTRIBUTING TO INDIVIDUAL DIFFERENCES Child Sex Developmental Considerations Behavioral Responses to Interparental Conflict (Stress) Characteristic Coping Behaviors and Cognitive Styles Neurobiological Responses to Interparental Conflict (Stress) Genetic Influences Interparental Conflict and the Broader Family and Community Context TRANSLATIONAL IMPLICATIONS Interventions Designed for Children Exposed to IPV Interventions to Reduce Interparental Conflict and Protect Child Well-being

Interventions Salient to Children's Exposure to Interparental Conflict Parenting Interventions that Reduce Interparental Conflict Mapping to the Organizational Framework and Future Directions DIRECTIONS FOR FUTURE RESEARCH Improving Theoretical Integration and Precision Translating Research into Practice CONCLUSION REFERENCES

Introduction and History Scientists and practitioners have long acknowledged that the quality of the parents' marital relationship affects children's development. For example, at the 1929 meeting of the American Orthopsychiatric Association, Towle (1931) discussed the importance of evaluating the marital relationship of potential foster parents, arguing “since a well-integrated marital situation affords the maximum possibility for well-adjusted children, it is essential that one understands the emotional life of the parents” (p. 281). A few years later, Hubbard and Adams (1936) concluded from an examination of case records from a child guidance clinic that parents' relationship to each other is among a handful of factors frequently found to have a bearing on the development of child problems. However, it was not until the 1970s that efforts to understand the link between marital functioning and child adjustment sharpened. This shift toward the systematic study of effects of marital functioning on children's development was prompted by several factors. Chief among these was the rising divorce rate in the United States, which stimulated research to better understand how parental divorce affects children. One conclusion that emerged from this research was that children's adjustment was determined more by the parents' conflict with one another than by the divorce itself (R. E. Emery, 1982). The burgeoning family therapy movement during the 1960s and 1970s also prompted interest in the link between marital and child functioning. A primary tenet of family systems theories is that individual family members are inextricably rooted in a larger family context that influences their thoughts and actions, and it was explicitly asserted that parents' marital dissatisfaction could stimulate and maintain adjustment problems in their children (Satir, 1964). Some speculation about this was quite bold. For example, published articles referenced the assertion that “whenever you have a disturbed child, you have a disturbed marriage” (Framo, 1965, p. 154), and this tenet was commonly accepted by adherents to family systems theory. There was also growing recognition among researchers outside of the family therapy movement (e.g., behavior therapists) that marital discord could adversely affect child adjustment and that clinicians might be able to enhance the effects of behavior therapy for children by attending to the parents' marital relationship (Oltmanns, Broderick, & O'Leary,

1977; Patterson, Cobb, & Ray, 1973). In addition, during this time, well-designed empirical studies were also beginning to identify marital discord as a potentially important determinant of child adjustment problems. For example, in a large community sample, Rutter (1971) found an unhappy marriage to increase risk for child conduct problems, even after controlling for other potential risk factors. During the latter part of the 1970s and the 1980s, research on marital functioning and child adjustment expanded considerably. Much of this research focused on answering some variation of the general question of whether there was a reliable association between parents' marital functioning and children's adjustment. Collectively, these early studies helped document the strength, robustness, and generalizability of the relation. It became clear that marital discord was indeed associated with child adjustment problems, in clinical as well as in nonclinical samples, across a very wide age range of children and across a variety of indices of marital functioning and child adjustment. Although the magnitude of the association between parents' marital functioning and child adjustment varied across studies, statistically significant associations, which were typically of small to medium effect size, emerged fairly regularly. It also started to become clear that certain dimensions of interparental conflict were more important than others in predicting child adjustment. These early studies laid the foundation for the development of two influential theories on the nature of the link between marital and child functioning: the cognitive-contextual framework (Grych & Fincham, 1990) and the emotional security hypothesis/theory (EST) (Davies & Cummings, 1994). Prior to the 1990s, much of the research on interparental conflict and child adjustment was guided by concepts from broad theories of family functioning or social behavior or only loosely tied to theory. These two theories were unique in that they specifically described processes by which parents' marital functioning might influence children's development and, in doing so, laid the groundwork for a new generation of research. They pushed the field to move beyond simply examining the extent to which marital and child functioning are related and to begin considering factors that help explain why this is so. In addition, because studies from the 1970s and 1980s also indicated that many children in maritally discordant families did not have significant adjustment difficulties, investigators also tried to identify those factors that seemed to shield children from the adverse effects of parental marital discord. Over two decades of research following the introduction of these two theories has now been amassed, and quite a bit of progress has been made toward understanding reasons why interparental conflict can lead to child problems and the conditions under which it might not. The next section of this chapter deconstructs the concept of interparental conflict, describing the dimensions of such conflict that, over time, can adversely affect children's social, behavioral, and emotional development. We then use these dimensions as a framework for reviewing the current state of knowledge on the association between interparental conflict and child adjustment problems. We follow the literature review with a presentation of several noteworthy explanations for the link between interparental conflict and child problems. In addition, we highlight key variables and processes implicated by this work and discuss how specific circumstances surrounding incidents of interparental conflict might influence the way

it relates to child adjustment problems. We follow this with a discussion of translational implications of this research and review examples of how the research has informed practice. In the final part of the chapter, we make suggestions for future directions for this research area. Throughout this chapter, we illustrate how different methodologies have been employed and how principles of developmental psychopathology have guided and should continue to guide research on this topic.

Interparental Conflict and Child Adjustment Problems Interparental conflict can be defined as opposition, disagreement, or a difference of opinion between parents; it is a normal part of adult relationships, and practically all children who have contact with their parents will be exposed to it (Papp, Cummings, & Goeke-Morey, 2002). Despite the simplicity of this definition, the phenomenon itself is complex, and the sources and manifestations of interparental conflict are many and varied. For example, a conflict may center around any number of topics, and the behaviors and emotions expressed during a single conflict incident can range widely in type and intensity. There is also variability in conflict resolution; some conflicts may be resolved quite easily with no untoward aftereffects, whereas others linger, fester, or seem to have been put to rest, only to arise again. These particular dimensions of conflict—topic, behavioral tactics, emotions, and resolution— provide the framework for our review. It is important to acknowledge at the outset that certain manifestations of these four dimensions are negative, but others may be benign or even positive and beneficial to children (Cummings & Davies, 2010). For example, parents whose disagreements are calm and respectful, and who come to a reasonable resolution of their differences, might provide their children with healthy models of conflict management and adaptive relationships. We make reference to some of the research that has examined positive aspects of interparental conflict in our review. Unfortunately, relatively little research has examined this side of the coin; more often the focus has been on understanding how conflict undermines child adjustment. In light of this fact, the bulk of our review focuses on the link between interparental conflict and problematic child adjustment. In most of the research on this topic, child adjustment problems are operationalized along the broad dimensions of externalizing problems, such as aggressive, oppositional, and delinquent behavior, or internalizing problems, such as symptoms of depression or anxiety. It should be acknowledged, however, that interparental conflict has also been linked to child adjustment problems that do not fit neatly into the categories of internalizing or externalizing problems. For example, exposure to interparental conflict has been linked to various aspects of child physical health (Nicolotti, El-Sheikh, & Whitson, 2003; Wood, Klebba, & Miller, 2000), cognitive and academic abilities (Ghazarian & Buehler, 2010; Harold, Aitken, & Shelton, 2007; Jouriles et al., 2008), and behavior in romantic relationships (Jouriles, Mueller, Rosenfield, McDonald, & Dodson, 2012; Kinsfogel & Grych, 2004). These findings are consistent with the developmental psychopathology principle of multifinality, which holds that a given risk factor (i.e., exposure to interparental conflict) may give rise to a variety of possible child outcomes. Although the focus in this chapter is on externalizing and internalizing

problems, other aspects of child adjustment are considered, as appropriate. Externalizing and internalizing problems typically are conceptualized as occurring along a continuum. It is not clear in many of the studies on this topic how many children in the samples were experiencing clinically significant levels of problems or outright psychological disorders, as defined by commonly used classification systems of mental disorders (e.g., Diagnostic and Statistical Manual of Mental Disorders). In fact, in many studies, it appears that most of the children in the samples were not experiencing clinically significant problems (as indexed by the standard measures that were used). In other words, it should be recognized that the variation in child adjustment problems accounted for in many studies appears to be primarily within the range of normal child adjustment. It is also noteworthy that many studies on this topic have evaluated children's immediate responses to specific incidents of adult conflict. In these studies, the investigators are measuring children's emotions (e.g., anger, sadness, fear) and behaviors (e.g., aggression) that occur immediately after a conflict incident. These are arguably symptoms of child distress, and not adjustment problems per se. However, there is evidence that children's immediate emotional and behavioral responses to interparental conflict, when collected across a number of incidents and aggregated, are associated with standard measures of child adjustment problems (Cummings, Goeke-Morey, & Papp, 2004; Goeke-Morey, Cummings, & Papp, 2007). Thus, this research is viewed as pertinent to the question of how interparental conflict influences children's adjustment.

Topic of the Conflict Compared to other dimensions of interparental conflict, research regarding the effect on children of the topic of the conflict is limited. However, results from studies using a variety of methods converge to suggest that conflict about child-related topics can be deleterious for children. For example, studies that make use of parents' reports on questionnaire measures of child-related interparental conflict indicate that such measures are associated with child externalizing and internalizing problems (e.g., Jouriles et al., 1991; Lee, Beauregard, & Bax, 2005; O'Leary & Vidair, 2005; Snyder, Klein, Gdowski, Faulstich, & LaCombe, 1988), and these associations emerge even after accounting for other dimensions of marital functioning, such as general marital adjustment. There is also experimental evidence that children are more likely to be immediately affected by interparental conflict that centers on a child-related topic as opposed to other topics. For example, in an analog study, Grych and Fincham (1993) had children listen to audio-recorded vignettes of conflict between a man and a woman. They asked the children to imagine that the conflict was between their parents and afterward had the children answer questions about their own affect, cognition, and potential coping strategies. Some of the conflicts focused on childrelated topics (homework time, who would take the child to an activity), whereas others focused on matters theoretically less related to the child (work, finances). Findings indicated that children reported more shame and fear of being drawn into the conflict when the conflict concerned the child. In another analog study, in which children watched videotaped vignettes

depicting conflict between a man and a woman (and were asked to imagine that the conflict was between their father and mother), unresolved child-related conflict resulted in more immediate anger and sadness than unresolved and resolved conflict focusing on non–childrelated topics (Koss et al., 2011). Other evidence for the immediate effects of child-related conflict on child behavior comes from a daily diary study conducted by Cummings and colleagues (2004). Mothers and fathers kept a home diary about incidents of marital conflict and children's aggressive responses to conflict over a 15-day period. Parents recorded different aspects of the conflict, including the topics discussed (child, marital, social, or work) as well as conflict tactics (behaviors exhibited to address the conflict) and emotions around the conflict. Findings indicated that interparental conflict about child and marital (e.g., intimacy, communication) topics related positively to child aggression in response to the conflict, whereas conflict about social and work topics did not. Moreover, the link between conflict about child-related topics and child aggression emerged even after accounting for conflict tactics and emotions surrounding the conflict.

Behavioral Tactics Research is accumulating to suggest that certain behavioral tactics used during parental disagreements are especially harmful to children. For example, studies in which intimate partner physical violence (IPV) is assessed indicate that children in families in which there has been a recent incident of IPV—common examples include pushes, grabs, and hits—are at greater risk for adjustment problems than children in families in which IPV has not occurred (see Kitzmann, Gaylord, Holt, & Kenny, 2003; Wolfe, Crooks, Lee, McIntyre-Smith, & Jaffe, 2003, for meta-analytic reviews of this literature). Moreover, this increased risk emerges for a variety of different types of adjustment problems (e.g., conduct-disordered behavior, emotional/ personality problems), even after accounting for other aspects of the parents' marital relationship, such as general marital discord (Jouriles, Murphy, & O'Leary, 1989). In addition, children's exposure to particular types of IPV, such as violence that carries a high risk for physical injury (violence involving knives and guns), even further elevates the risk for child adjustment problems (Jouriles et al., 1998; Spaccarelli, Coatsworth, & Bowden, 1995). There is also experimental evidence indicating that children react more negatively to conflict involving IPV than to conflict involving other behavioral tactics. Specifically, evidence from studies employing vignette methods similar to those of the Grych and Fincham (1993) study described in the “Topic of the Conflict” section indicates that conflict involving physical aggression between adults elicits more distress and negative emotion from children than does verbal conflict (e.g., Cummings, Vogel, Cummings, & El-Sheikh, 1989; Cummings, ZahnWaxler, & Radke-Yarrow, 1981). In addition to physical violence, psychologically and emotionally abusive behaviors, such as insults, verbal threats, and aggression toward objects, are also associated with child adjustment problems. For example, Jouriles, Norwood, McDonald, Vincent, and Mahoney (1996) found that parents' reports of psychological aggression relate to child adjustment problems, even after accounting for acts of physical IPV. Other investigators have documented

similar associations, highlighting the potentially unique contribution of these tactics in understanding the link between interparental discord and child adjustment problems (Clarke et al., 2007). Experimental evidence also suggests that interparental psychological and emotional abuse can cause immediate distress in children that is sometimes just as distressing as physical violence (Laumakis, Margolin, & John, 1998). The two behavioral tactics just discussed (physical violence and psychologically or emotionally abusive behavior) are overt forms of conflict expression. Such behavior appears to be harmful to children, regardless of which parent engages in it (see Sturge-Apple, Skibo, & Davies, 2012, for a review that addresses this issue), and it is most harmful if both parents engage in it, as compared to only one parent (McDonald, Jouriles, Tart, & Minze, 2009). There also is some indication that more subtle forms of conflict expression, such as withdrawal or detachment (e.g., one parent ceasing to attend to what the other parent is saying), particularly in combination with overt forms of conflict expression, are harmful to children. For example, Katz and Woodin (2002) classified couples on the basis of their conflict interactions and found that child adjustment problems were most likely when the parents engaged in both attacking and withdrawing tactics during conflict. Goeke-Morey, Cummings, and Papp (2007) found that interparental conflicts that ended with one of the parents withdrawing from the interaction were associated with greater child behavioral dysregulation (misbehavior, aggression) in response to the conflict. Related to these findings, results from experimental analog studies indicate that children are sensitive to adult withdrawal and nonverbal expressions of anger (e.g., angry glares) during adult conflicts, and they respond to these behaviors with negative emotions (Cummings et al., 1989; Goeke-Morey et al., 2007). Although relatively little research has evaluated effects of behavioral tactics that might be conceptualized as positive, the results of several studies suggest that this might be a fruitful avenue for future research. For example, Davies, Martin, and Cicchetti (2012) created a latent variable of constructive marital conflict, which included parents' reports of responding positively toward one another during challenging/stressful times, as well as cooperation and satisfactory conflict resolution. This latent variable was negatively associated with later child adjustment problems, even after accounting for destructive interparental conflict. McCoy, Cummings, and Davies (2009) created a similar latent variable, which included data derived from observations of parents during a conflict interaction (e.g., problem solving, displays of affection) as well as parents' reports of cooperation and satisfactory resolution of their conflicts in general. Although constructive marital conflict was not directly related to later prosocial child behavior, indirect relations emerged in intervening variable models. Frosch and Mangelsdorf (2001) created a composite code, labeled positive engagement, for the coding of a nonconflictual marital interaction task. This code included behaviors such as initiating conversation and demonstrating involvement in the task, cooperation, and the extent to which spouses affirmed each others' contributions. In this study, positive engagement was negatively associated with observer ratings of child externalizing problems during a home visit. Moreover, positive engagement was negatively associated with child externalizing problems even after controlling for marital conflict. It is important to highlight that in the few studies considering both positive and negative

conflict tactics (or constructive and destructive interparental conflict), there was evidence that both contribute uniquely to child adjustment problems (Davies et al., 2012; Frosch & Mangelsdorf, 2001). In other words, it appears that it is not just the presence of destructive interparental conflict that increases risk for child adjustment problems, but also the absence of constructive or positive interparental interactions. Although some investigators have explored whether positive conflict tactics can mitigate the effects of negative conflict tactics on child adjustment problems, the results from a couple of well-designed studies suggest that positive and negative conflict tactics do not interact with one another in the prediction of child adjustment problems (Davies et al., 2012; Frosch & Mangelsdorf, 2001). We revisit this issue again in the section on conflict resolution.

Emotions The emotions experienced and the behaviors expressed during conflict may be inextricably linked, although they are arguably separate phenomena. This link, however, may partly be an artifact of the methods used to assess interparental conflict. Specifically, some questionnaire items used in studies linking interparental conflict with child adjustment problems combine angry and hostile emotions with violent and abusive behaviors; the same could be said for the conflict vignettes that are used in experimental studies. Studies attempting to isolate or disentangle negative affect (hostility, anger) from destructive tactics during conflict suggest that both contribute independently to children's aggressive behavior (Cummings et al., 2004). Very little research has evaluated the role of positive emotion expressed during interparental conflict. In one of the few studies on this topic, Goeke-Morey, Cummings, and Papp (2007) found that parental positive emotions expressed during conflict endings were positively associated with child positive emotions and negatively associated with child negative emotions (anger, sadness, fear) immediately following the conflicts. Similarly, Cummings and colleagues (2004) found that positive emotions expressed by either parent during conflict were associated with a lower likelihood of child aggression in response to the conflict. However, there was an important caveat to this latter finding. Conflicts about child-related topics were still positively associated with aggressive child responses, regardless of the nature of the emotions expressed during the conflict.

Resolution of Conflict Results from a series of experimental analog studies indicate that poorly resolved conflicts, such as those that end with continued hostility or “the silent treatment,” elicit negative immediate reactions from children (e.g., Cummings, Vogel, Cummings, & El-Sheikh, 1989; Cummings, Ballard, El-Sheikh, & Lake, 1991). By comparison, resolved conflicts are associated with more positive immediate reactions from children, and there is some evidence that certain tactics and emotions displayed during the resolution period can even ameliorate some of the immediate negative effects of hostile conflicts on children. Specifically, children seem to respond positively to resolution that involves positive affect and compromise (GoekeMorey et al., 2007; Koss et al., 2011; Shifflett-Simpson & Cummings, 1996). Apology appears beneficial when coupled with positive affect, but not when coupled with sadness or anger

(Goeke-Morey et al., 2007). Interestingly, the effects of resolution are robust, emerging across many different types of conflict scenarios, including those that include physical aggression (Cummings et al., 1991; El-Sheikh, Cummings, & Reiter, 1996). In one of the few field studies examining couples' resolution of interparental conflict and child adjustment problems, GoekeMorey and colleagues (2007) found that parent ratings of the degree to which the problem had been resolved were negatively associated with child negative emotions (anger, sadness, fear) and problematic behaviors (misbehavior, aggression) immediately following the conflicts. Similarly, Kerig (1996) found that a questionnaire measure reflecting the degree to which positive or negative affect dominates the resolution of conflict correlated in the expected direction with several indices of child adjustment.

Summary on Dimensions of Interparental Conflict In sum, empirical evidence derived from correlational and experimental research converges to suggest that certain aspects of interparental conflict are more destructive to children than others. In particular, conflict about child-related topics and conflict that includes physical violence or psychologically abusive behavior appears to be especially detrimental to children. There is also some evidence suggesting that parental withdrawal and detachment during conflict episodes is harmful. Heightened negative emotions during conflict interactions contribute to the adverse effects of interparental discord on children. Positive approaches to interparental conflict, such as parents cooperating with one another, contribute to positive child adjustment. In addition, the outcome and immediate aftermath of conflict situations are important, with conflicts that are resolved in a positive manner (this might include compromise, apology, and the expression of positive emotions) being less harmful to children.

A Note about Frequency Conceptually, research on interparental conflict and child adjustment is concerned with the effects of interparental conflict writ large. Only in extreme instances would one or two isolated incidents of conflict be expected to have marked and lasting effects on a child's development. Thus, in addition to considering the dimensions of conflict just reviewed (topic, behavioral tactics, emotions, and resolution), understanding how interparental conflict affects children's adjustment also demands considering the frequency of the conflict. Generally speaking, a child's history of exposure to destructive forms of interparental conflict influences how that child will respond to a specific conflict incident, and the frequency of children's exposure to destructive forms of interparental conflict correlates positively with the degree to which children experience adjustment problems. That is, the more frequently parents have disagreements with child-related content, the more likely the children will experience adjustment problems. Similarly, the more frequently physical IPV occurs, the more likely the children will experience adjustment problems, or the more frequently psychologically and emotionally abusive behaviors occur, the more likely there will be child problems. In other words, this frequency dimension cuts across each of the individual aspects of interparental conflict already reviewed. Over time, however, the frequency of interparental conflict is likely to vary. There may be

some periods in which interparental conflict is relatively frequent and some in which it is not. Although studies addressing this issue are sparse, there is evidence that child problems change with changes in frequency. For example, in a three-wave longitudinal study, elevations in interparental conflict over time corresponded to elevations in child externalizing problems during this same time period (Kouros, Cummings, & Davies, 2010). Similarly, changes in marital distress were positively correlated with changes in adolescents' adjustment (anxiety, depression, delinquency and hostility) over a 5-year period (Cui, Conger, & Lorenz, 2005). A related consideration pertains to the trajectories of specific types of interparental conflict over the course of time, and there is some indication that the trajectory of children's exposure to interparental conflict is important. For example, within a sample of school-age children who had recently been exposed to IPV, increases in children's exposure to IPV over time were related to increases in childhood depression (Kennedy, Bybee, Sullivan, & Greeson, 2010). Similarly, more consistent or chronic exposure to IPV predicts the development of child internalizing and externalizing problems, more so than more sporadic exposure (e.g., C. R. Emery, 2009; Martinez-Torteya, Bogat, von Eye, & Levendosky, 2009).

Developmental Considerations The bulk of the research linking interparental conflict and child adjustment problems has been conducted with families of preschoolers, school-age children, and adolescents rather than with infants. Among the studies with infants, however, positive associations have been demonstrated between interparental conflict and general emotional and behavioral distress displayed by infants, insecure/disorganized attachment behaviors, and temperament characteristics associated with poor adjustment outcomes (e.g., Du Rocher Schudlich, White, Fleischhauer, & Fitzgerald, 2011; Owen & Cox, 1997; Pauli-Pott & Beckmann, 2007). For example, Du Rocher Schudlich and colleagues (2011) had parents discuss an area of disagreement in the presence of their infant (ages 6–14.5 months). Destructive conflict—a composite variable that included contempt, defensiveness, demand, and anger—was related to infant distress, dysregulation, and physical frustration. Infants appear to be sensitive to positive aspects of interparental conflict as well. In this same study, constructive conflict tactics, which included problem solving, support, and positive emotion, were associated with less infant physical and emotional frustration and dysregulation. Thus, interparental conflict is associated with indicators of child adjustment for children of all ages. It might also be argued that certain forms of interparental conflict influence child functioning, even prior to the birth of the child. For example, although empirical data are limited, women who experience IPV during pregnancy, compared to those who do not, are at increased risk for giving birth to a low-birthweight infant (Murphy, Schei, Myhr, & DuMont, 2001). Similarly, IPV during pregnancy has been linked to a weaker maternal-fetal bond and less secure maternal representations of caregiving (Huth-Bocks, Levendosky, Bogat, & von Eye, 2004; Zeitlin, Dhanjal, & Colmsee, 1999); there is also speculation that IPV during pregnancy can have lasting effects on child stress-regulatory systems (Graham, Kim, & Fisher, 2012; Levendosky, Leahy, Bogat, Davidson, von Eye, 2006). Consistent with these findings, IPV during pregnancy has been linked with child adjustment problems during the preschool

years (Flach et al., 2011; Levendosky et al., 2006). Moreover, investigators have documented associations between mothers' lifetime experiences of IPV and both fetal loss and infant death (Ahmed, Koenig, & Stephenson, 2006). It is not clear yet if there is a “sensitive period” over the course of development, during which exposure to interparental conflict is more strongly associated with child adjustment problems, as compared to other time periods. In fact, this question has received surprisingly little attention in the research literature. Yates and colleagues (2003) hypothesized that early exposures to IPV might have an especially strong and enduring effect on subsequent child development, as compared to later exposures. This is because it is during these early years that children are developing the ability to regulate their emotions and behaviors. In other words, children are developing the foundation for adapting to their environment, both in the present and in the future. In addition, during these early years, the child's foundation is believed to be especially sensitive to extreme environmental influences, such as IPV. This is not to say that later exposures to IPV are unimportant. Rather, later exposures are not expected to compromise a child's core self-regulation and adaptive stress-response capacities in the same way that early exposures might. Yates and colleagues (2003) tested their hypothesis through the use of a longitudinal data set that followed children from preschool to 16 years. These investigators estimated the degree of male-to-female partner violence (referred to as IPV) based on interview data collected from mothers and items from Life Events Scales. These estimates were provided for two time periods in the children's lives: preschool years: 18–64 months; middle childhood: grades 1–3. Children's exposure to IPV during the preschool years made a unique contribution to adolescent (age 16) externalizing problems, above and beyond a number of control variables (e.g., socioeconomic status [SES], life stress) as well as exposure to IPV during the middle childhood years. In fact, IPV during the middle childhood years was not a predictor of externalizing problems at age 16. Similar findings emerged for IPV during the preschool years and child internalizing problems when the sample was limited to girls. Conclusions drawn from reviews of the literature suggest the magnitude of the relations between interparental conflict and child adjustment problems are fairly consistent across age groups. For example, Buehler and colleagues (1997) found that the effect size for the relation between interparental conflict and child adjustment (.32) did not differ based on the age or age range of children included in studies. Similarly, Evans, Davies, and DiLillo (2008) found medium effect sizes for IPV and child internalizing (effect size range: .47–.51) and externalizing (effect size range: .40–.49) problems, which did not differ based on whether children in the studies were zero to 5, 6 to 12, or 13 to 19 years old. These reviews, however, examined primarily cross-sectional studies on the relation between interparental conflict and child adjustment problems. In sum, it seems clear that children's exposure to interparental conflict, at any age, can influence child adjustment problems. It is not yet clear, however, if there are specific age periods in which children's exposure to interparental conflict is especially detrimental to children (i.e., more detrimental than other age periods). Another developmental question pertains to how enduring or long-lasting the link is between

interparental conflict and child adjustment problems. It is clear from experimental research that interparental conflict can cause immediate distress among witnessing children. It is also clear from the extant correlational and longitudinal research that this is not just an immediate, shortterm effect. Specifically, prospective relations between exposure to interparental conflict and adjustment difficulties have been documented over time periods extending in excess of 10 years (Linder & Collins, 2005; Yates, Dodds, Sroufe, & Egeland, 2003). Yet there is also evidence that some of the adverse effects of witnessing IPV dissipate during the adult years. Specifically, longitudinal studies on the relation between adolescents' exposure to IPV and their involvement in violent romantic relationships suggests a relatively strong association during adolescence and even early adulthood; however, this association appears to dissipate over the course of adulthood (Smith, Ireland, Park, Elwyn, & Thornberry, 2011). One approach to this question about the enduring effects of interparental conflict, which acknowledges the complexity in which such conflict might exert an influence on development over time, is the construction and evaluation of developmental cascade models, which propose how interparental conflict might interfere with stage-salient tasks, thereby increasing the risk for future adjustment problems (Cicchetti, 1993). Kouros, Cummings, and Davies (2010) tested a developmental cascade model and found that increasing interparental conflict in early childhood (kindergarten to second grade) was related to poor social competence in preadolescence (age 12) via externalizing problems in early childhood. The results suggested that interparental conflict may interfere with children's stage-salient tasks in early childhood, resulting in increased externalizing problems. Externalizing problems in early childhood, in turn, interfere with stage-salient tasks needed to promote optimal social competence and prosocial behavior. Similarly, Davies, Manning, and Cicchetti (2013) found evidence for a developmental cascade model in which preschoolers' exposure to interparental conflict is related to behavior problems 2 years later by interfering with the stage-salient tasks of regulating emotions, establishing autonomy, and problem solving. Additional developmental cascade models are reviewed in the section titled “Theory Explaining the Link between Interparental Conflict and Child Adjustment.” Developmental cascade models that select stagesalient tasks specific to the developmental period studied offer promise for understanding developmental pathways that lead to different child adjustment outcomes across individuals and across time.

Prevalence and Degree of Children's Exposure to Interparental Conflict It is impossible to say with certainty exactly how many children are exposed to destructive forms of interparental conflict. However, one might get a sense of this from statistics on the prevalence of children's exposure to IPV, parental divorce, and marital distress. Epidemiological data suggest that a large number of children in the United States are exposed to IPV and psychological abuse. Although annual prevalence estimates of children's exposure to IPV (e.g., pushing, grabbing, slapping) vary widely from study to study (e.g., Bair-Merritt et al., 2008; Hamby, Finkelhor, Turner, & Ormrod, 2011; C. G. Moore, Probst, Tompkins, Cuffe, & Martin, 2007), IPV has been estimated to occur in as many as 30% of dual-parent

households with children in the United States (McDonald, Jouriles, Ramisetty-Mikler, Caetano, & Green, 2006), with severe physical IPV (kicks, punches, threat or use of weapons) estimated to occur in as many as 13% of dual-parent households with children (McDonald, Jouriles, Ramisetty-Mikler et al., 2006). Psychological abuse between parents or caregivers (e.g., threatening, insulting, or attempting to control a partner) appears to be even more common; a population-based survey of mothers in North and South Carolina indicated that psychological abuse occurs in the homes of nearly 50% of children over the course of a year (Chang, Theodore, Martin, & Runyan, 2008). General marital distress and divorce are more indirect markers of children's exposure to interparental conflict, as compared to the statistics regarding children's exposure to IPV. Nevertheless, statistics about their prevalence offer a general sense of the degree to which children are exposed to potentially harmful forms of interparental conflict. Among intact couples, close to one-third report having distressed marital relationships at any given stage of the relationship (Whisman, Beach, & Snyder, 2008). Moreover, there is some indication that marital distress might be slightly more prevalent during the years in which couples are raising young children. Approximately 40% of American children experience divorce during childhood, with an estimated 5% to 25% of divorces involving high levels of interparental conflict (Lamb, Sternberg, & Thompson, 1997; Maccoby & Mnookin, 1992). Thus, it is safe to conclude from existing studies that, each year, parents of millions of American children engage in forms of interparental conflict that have been found to be harmful to children. Yet the data just reported provide only an idea of how many children live in families in which destructive interparental conflict occurs. At a micro-model level, a key contextual variable is the extent to which children are actually exposed to the conflict. For example, are children present when the conflict occurs, and do they directly witness it? This is in contrast to other forms of exposure to interparental conflict, which include living in a household in which it occurs but being unaware of it, or hearing about it from siblings or the parents themselves. Results from a handful of studies suggest that children are present during over a third of their parents' destructive interparental conflict episodes. Specifically, in a study examining policereported incidents of IPV, Fantuzzo and Fusco (2007) found that children were present during 43% of the violent incidents, and when they were present, they directly observed the violence 80% of the time. Moreover, the authors also found that children were disproportionately exposed to what they described as the most unstable and dangerous types of IPV, which included weapon use and mutual assault. Papp, Cummings, and Goeke-Morey (2002) trained parents to keep diaries of their conflicts over a 15-day period and found that approximately one-third of the conflicts occurred in the children's presence. Interestingly, conflicts in which children were present, compared to those in which they were absent, were more often about child-related issues and more likely to include negative emotions and destructive conflict tactics.

Theory Explaining the Link between Interparental Conflict and Child Adjustment

Before the 1980s, much of the research on interparental conflict and child adjustment was guided only loosely by theory and was designed simply to examine whether indicators of a troubled marriage (e.g., parent reports of poor marital adjustment) were associated with various kinds of child adjustment problems. In the late 1970s and 1980s, researchers began to apply broad theories of social behavior, such as Bandura's (1973; 1986) social cognitive theory (earlier versions known as social learning theory), to the topic of interparental conflict and child adjustment. Social cognitive theory pointed to a troubled marriage as a risk factor for child adjustment problems, and, more important, it highlighted processes by which a troubled marriage might increase risk (via children observing and modeling/learning from particular expressions of interparental conflict). This theory sparked some of the research on different manifestations of interparental conflict and their relation to child adjustment problems. For example, some of the studies in the late 1970s and 1980s were designed to evaluate whether the particular types or dimensions of interparental conflict that social cognitive theory suggested ought to be important (e.g., physical aggression, overt hostility) contributed to child adjustment problems after accounting for more diffuse measures of marital problems, such as poor global marital adjustment (e.g., Jouriles, Murphy, & O'Leary, 1989; Porter & O'Leary, 1980). Although there was recognition during the late 1970s and 1980s that social cognitive theory might help explain the link between interparental conflict and child adjustment problems, researchers were also invoking other theories and ideas to help explicate documented relations. For example, in the 1980s, E. Mark Cummings and his colleagues conducted a number of studies focusing on very young children's (e.g., toddler-aged) immediate reactions to inter-adult anger and aggression. In this research, it was noted that the children did not appear to be learning aggression from these adult interactions, because their responses did not seem to precisely mimic or imitate what the adults did. Rather, the children seemed to be experiencing heightened arousal and negative emotions from these adult interactions, which transferred to other interactions that the children were having subsequent to the conflict incident. In short, children's emotional reactivity to the adult conflict was thought to be key (in contrast to learning per se), and contagion of emotion and excitation transfer theories were being used to help explain empirical observations from this early research (Cummings, Iannotti, & Zahn-Waxler, 1985). Theory may have played a role in guiding some of the early research on interparental conflict and child adjustment, but theoretical processes invoked to explain why interparental conflict might lead to child problems were rarely tested in these early studies. Rather, the focus was on developing a better understanding of the parameters of the relation between interparental conflict and child adjustment problems. In the 1990s, theories specific to the link between interparental conflict and child adjustment began to appear in the scientific literature. Two of the more prominent theories were Grych and Fincham's (1990) cognitive-contextual framework and Davies and Cummings's (1994) EST. These two theories helped prompt and guide the development of more process-oriented studies designed specifically to explain how interparental conflict might lead to child adjustment problems. Many of the studies conducted during the 1990s and afterward were direct tests of components of the cognitive-contextual

framework or EST. In addition, new theoretical explanations for the link between interparental conflict and child adjustment problems began to arise and other preexisting ones were revisited during these years. In this section of the chapter, we describe some of the theories and explanations for the link between interparental conflict and child adjustment problems in more detail. Following the lead of other investigators (Davies & Cummings, 2006), we organize this section by distinguishing between direct and indirect theorized effects of interparental conflict on the child. The “direct-pathway” or “direct-effects” models suggest that children's exposure to destructive interparental conflict triggers processes within the child (e.g., cognitions, emotions, biological factors) that increase their risk for adjustment problems. In contrast, the “indirect-pathway” or “indirect-effects” models suggest that interparental conflict influences variables outside of the child (parenting behavior) that lead to child adjustment problems. It is important to acknowledge several points before proceeding. First, many explanations of the link between interparental conflict and child adjustment have been advanced. Our review is not intended to be exhaustive. Rather, we cover what we believe to be the most influential theories on this topic as well as several other interesting theories that have received relatively little research attention but that help give a sense of the breadth of explanations that have been offered to explain this link. Second, several of the theories on the link between interparental conflict and child adjustment are quite elaborate; some have had entire books devoted to them (e.g., EST; Cummings & Davies, 2010). We review what we believe are the key ideas and concepts of these theories because a comprehensive review of each theory is beyond the scope of this chapter. Also, several of the more comprehensive theories include both direct and indirect pathways by which interparental conflict influences child adjustment problems. For example, EST (Davies & Cummings, 1994) proposes that children's exposure to destructive interparental conflict triggers processes within the child that can lead to child adjustment problems; in addition, EST accepts the notion that interparental conflict influences whether, and how, children use their parents as a source of support and protection, which can also lead to child adjustment problems. By organizing our review into direct and indirect pathways, some of the complexity and nuance of the more comprehensive theoretical explanations is unfortunately lost. Third, the vast majority of the research on the topic of interparental conflict and child adjustment problems has been conducted from the perspective that interparental conflict influences child behavior. Child-effect hypotheses are rarely examined in this research literature. Yet early studies offered some acknowledgment that child behavior problems influence the nature of the interparental conflict. For example, interparental conflict on childrelated topics has been suggested to be more common in families in which the children are exhibiting problematic behaviors, because these behaviors might prompt parents to discuss, and perhaps disagree about, how to address them (Jouriles et al., 1991). More recently, investigators have begun to examine how certain child responses to interparental conflict (e.g., helping behavior performed by the child with the intent to diminish interparental conflict) might influence subsequent interparental interactions (e.g., Schermerhorn, Cummings, DeCarlo, & Davies, 2007). In addition, investigators have also begun to examine how child problems

themselves might lead to more severe forms of marital discord and dissolution (e.g., Wymbs et al., 2008). Our review is limited to theories and ideas on how interparental conflict influences children, but we acknowledge that the effects likely go both ways. Fourth, it is now generally accepted that “any single theory is inadequate in capturing the reality of children's lives” in highly conflictual or violent families (Jaffe, Wolfe, & Campbell, 2012, p. 17). Researchers recognize that certain aspects of the various theories on the link between interparental conflict and child adjustment are complementary, or at least not in opposition to one another (e.g., Davies & Cummings, 1998). Indeed, many possess similar concepts or ideas but label and/or package them differently. For example, children's “internal representations” (e.g., EST) and “cognitive scripts” (e.g., social cognitive theories) of interparental conflict essentially refer to the same general concept: structures/processes within children that probabilistically govern what they anticipate will happen once a parental conflict has started.

Direct-Effects Models Social Learning and Social Cognitive Theory Social learning and social cognitive theories suggest that children learn from observing others, and parents are powerful models for children's observational learning (e.g., Bandura, 1973; 1986). In the case of interparental conflict, parents model ways of addressing and handling interpersonal conflict. Some parents model angry and hostile conflict behaviors, failing to provide their children with respectful and productive models (Margolin, 1981). Through observing their parents' conflicts over time, children learn or acquire, among other things: (1) a repertoire of behavioral tactics to use in conflict situations, (2) cognitions about the effectiveness of those tactics in achieving desired outcomes, (3) cognitive scripts for how conflict situations play out, and (4) general beliefs about the acceptability and justifiability of certain types of behavioral tactics, including which are condoned or considered permissible. Children who repeatedly observe destructive conflict (e.g., angry and hostile parental interactions) presumably acquire behavioral tactics and cognitions that contribute to the development of angry and hostile social interaction patterns. Social cognitive theories on the development and perpetration of aggressive and violent behavior sometimes distinguish between the development of explicit and implicit cognitions (e.g., C. A. Anderson & Bushman, 2002; Crick & Dodge, 1994; Jouriles, McDonald, Mueller, & Grych, 2012). Explicit cognitions are the articulated ideas of conscious and effortful thought. An example is the belief that aggressive behavior is justified under certain circumstances or that it facilitates the achievement of desired outcomes. Implicit cognitions, in contrast, can be unarticulated, automatic, and outside of awareness. Knowledge structures that increase access to aggressive thoughts are an example of an implicit cognitive mechanism that heightens the likelihood of aggressive behavior. More precisely, increased access to aggressive thoughts is hypothesized to increase the likelihood that ambiguous situations will be interpreted in a hostile manner. Social cognitive theories posit that explicit and implicit cognitions that facilitate the likelihood of aggressive behavior develop as a result of social experiences,

including the observation of destructive interparental conflict, and, once developed, these two different types of cognitions operate jointly to shape future behavior. Dodge and colleagues' social information processing model has been especially influential in enhancing the understanding of child and adolescent aggressive behavior (Crick & Dodge, 1994; Dodge, 1986). It has also been applied to the topic of interparental conflict and child adjustment. This model proposes four sequential processes that culminate in a response to a social situation: (1) encoding, which is attending to the external and internal cues in the situation; (2) interpretation, which involves formulating attributions regarding others' intent; (3) response generation, which is generating possible responses to the situation; and (4) response evaluation, which is evaluating the possible responses based on the child's goals, anticipated results, and self-efficacy for enacting the response. Similar to other social cognitive theories, a child or adolescent's information-processing style is hypothesized to develop as a result of their social experiences, including the observation of interparental conflict. Dodge's model suggests that by observing how their parents' conflicts emerge and are handled, children may note and internalize situational cues that bear on conflict, make inferences about motives pertaining to conflict, gain a specific repertoire of possible responses to conflict, and evaluate the effectiveness and consequences of those responses. Once developed and integrated, this information-processing style is hypothesized to influence children's future social interactions. Quite a bit of research examining the link between interparental conflict and child adjustment, especially child externalizing problems or aggressive behavior, is consistent with tenets of social cognitive theory. Most studies directly evaluating processes derived from social cognitive theories find that children exposed to destructive forms of interparental conflict are more likely to have cognitions that facilitate aggressive behavior, such as attitudes more accepting of violence or a hostile attribution bias (Calvete & Orue, 2011; Marcus, Lindahl, & Malik, 2001). These cognitions, in turn, have been found to mediate the association between interparental conflict and various types of child adjustment problems, especially those that involve aggressive behavior. For example, a number of researchers have found key social cognitive processes to mediate associations between different forms of interparental conflict (especially IPV) and violence in romantic relationships (Fite et al., 2008; Kinsfogel & Grych, 2004; Riggs & O'Leary, 1996). It should be noted, however, that most of the research evaluating associations between interparental conflict and the cognitive processes proposed by social cognitive theory focuses on older school-age children and adolescents. Before concluding this section, it is important to acknowledge the role of reinforcement in social learning theories. Both positive (providing a positive stimulus) and negative (removing a negative stimulus) reinforcement of behaviors increase the likelihood that those behaviors will be repeated. Through observing their parents' conflicts and experiencing their own conflicts with others, children learn what is likely to be reinforced, thus shaping their own behavior and cognitions, in conflict situations and beyond. That is, through reinforcement (vicarious and direct), children form expectancies about the utility and consequences of certain behavioral tactics and develop cognitive scripts for how conflict situations are likely to play out. These expectancies and scripts, in turn, guide future behavior.

Emery and colleagues (R. E. Emery, 1989; R. E. Emery & Kitzmann, 1995) describe a model of children's immediate reactions to interparental conflict, which illustrates how negative reinforcement might shape children's behavior in conflict situations and has implications for the development of more general child adjustment problems. In this model, interparental conflict is viewed as an aversive event for children that can prompt a variety of negative cognitive, affective, and behavioral reactions, some of which might be construed as antisocial (yelling at or arguing with a parent). An antisocial child reaction that is sufficient to prompt the parents to stop their immediate conflict negatively reinforces that child reaction. That is, the aversive event, interparental conflict, is removed; thus, according to the principle of negative reinforcement, the child reaction that prompted the removal of the aversive event (displays of fear or anger, yelling at or arguing with a parent) is likely to be repeated. The model further proposes that, over time, through repetition of this family dynamic, this pattern of cognitive, affective, and behavioral reactions to conflict becomes internalized and ingrained in the child. Consistent with Emery's theory, research on children's immediate responses to conflict suggests that exposure to destructive conflict is aversive for children, prompting reactions of fear and anger (see Cummings & Cummings, 1988; Cummings & Davies, 2010; Davies & Cummings, 1994, for reviews). In fact, most children have an adverse emotional response to their parents' arguments (Peterson & Zill, 1986). In addition, there is quite a bit of evidence that destructive forms of interparental conflict sometimes prompt behavioral reactions from children that might be interpreted as aggressive or oppositional, such as yelling and arguing with a parent (Jenkins, Smith, & Graham, 1989; Shelton & Harold, 2008). These immediate affective and behavioral reactions to interparental conflict have been linked to more general measures of child behavior problems (e.g., Cummings et al., 1989). However, it is not entirely clear from the available data whether negative reinforcement is the key mechanism operating to link children's immediate reactions to interparental conflict with more general measures of their adjustment. Cognitive-Contextual Framework According to Grych and Fincham's (1990) cognitive-contextual framework, when children witness an incident of interparental conflict, they attempt to derive meaning from it. They do so by appraising the situation and attempting to understand what is happening and how it might affect them and their family (threat appraisals), as well as why it is happening (blame appraisals) and what they can do in response (coping appraisals). These appraisals are posited to lead to behavioral responses to the immediate situation (coping behaviors). In figures illustrating the framework (e.g., Grych & Cardoza-Fernandes, 2001; Grych & Fincham, 1990), appraisal processes and behavior are depicted in a sequence of three steps or stages, with threat appraisals (primary processing of the conflict) preceding blame and coping appraisals (secondary processing of the conflict), which, in turn, lead to a behavioral response (coping behavior). However, Grych and colleagues acknowledge that these steps can occur simultaneously and that their influence on each other can be reciprocal. Children's threat and blame appraisals are held to be especially important. Although the model is called “cognitive”-contextual, these appraisals are actually conceptualized to include both cognitive

and affective processes. For example, threat appraisals include thoughts about danger (e.g., a parent might be harmed or the family might split up) and the emotion of fear, whereas blame appraisals include thoughts about responsibility and, with self-blame, the emotion of guilt. The cognitive-contextual framework offers hypotheses about how specific aspects of interparental conflict and the family context might influence children's threat and blame appraisals (Grych, 1998; Grych & Fincham, 1990). For example, when interparental conflict is about a child-related issue, as compared to something else (e.g., parents' work), it is hypothesized that children are more likely to blame themselves for the conflict. It is also hypothesized that conflict about child-related issues is likely to be more threatening to children, because of the potential spillover into parent-child interactions and negative consequences for the children. As interparental conflict becomes more hostile and angry and includes physically aggressive and emotionally abusive behaviors, it is hypothesized that children are more likely to view the conflict as threatening. Children's prior exposure to interparental conflict is hypothesized to influence appraisals as well, because these previous experiences have presumably shaped children's expectations about the course of future conflicts. Prior exposure to IPV, for example, is expected to sensitize children to the possibility that future conflicts might also culminate in violence. The original formulation of the cognitive-contextual framework (Grych & Fincham, 1990) focused on children's immediate responses to interparental conflict—that is, their appraisals and coping behaviors in response to an incident of conflict. Yet they also offered some hints about how particular appraisals and coping behaviors in response to a conflict incident might lead to maladaptive behavior more broadly. For example, they made reference to R. E. Emery's (1989) model described earlier in the section “Social Learning and Social Cognitive Theory” as a process that might maintain children's maladaptive cognitive, affective, and behavioral responses to interparental conflict. In later writings, Grych and colleagues expanded on how these immediate appraisals and coping behaviors might lead to child adjustment problems more broadly. For example, Grych, Seid, and Fincham (1992) proposed that children who frequently appraise interparental conflict as threatening may develop symptoms of anxiety disorders, particularly if the conflict occurs very frequently; similarly, children who frequently appraise interparental conflict as their fault may develop symptoms of depressive disorders. Along this same line, children who appraise interparental conflict as both threatening and their fault, and who are unable to stop it from occurring, may develop a sense of hopelessness that leads to a variety of internalizing problems (Grych, Fincham, Jouriles, & McDonald, 2000). Many studies find that as children feel increasingly threatened by or at fault for interparental conflict, their risk for adjustment problems increases. In fact, a robust empirical literature has linked children's threat and self-blame appraisals to their internalizing problems (Rhoades, 2008). Although there is some evidence that these appraisals are also linked to externalizing problems, the findings are much more consistent with internalizing problems (Fosco et al., 2007; Rhoades, 2008). It should also be acknowledged here that there is quite a bit of evidence suggesting certain child coping behaviors in response to interparental conflict (particularly children attempting to interrupt parental conflict in an aggressive or oppositional manner) are

consistently linked with broad measures of children's internalizing and externalizing problems (Rhoades, 2008). Moreover, the results of a handful of studies have now accumulated to suggest children's threat appraisals, self-blame appraisals, or both mediate effects of interparental conflict on child adjustment problems (e.g., Fosco & Grych, 2008; Gerard, Buehler, Franck, & Anderson, 2005; Grych et al., 2000; Grych, Harold, & Miles, 2003; Kim, Jackson, Conrad, & Hunter, 2008; Mann & Gilliom, 2004). Most of this research has focused on school-age children and young adolescents, but the theorized processes linking interparental conflict with child adjustment are likely to apply to younger children as well, with the caveat that their appraisals are likely to be less complex and more concrete than those of older children (McDonald & Grych, 2006). Consistent with the cognitive-contextual framework, the family context in which the interparental conflict occurs influences how interparental conflict relates to children's threat and blame appraisals. Deboard-Lucas, Fosco, Raynor, and Grych (2010) demonstrated how several different family variables moderate relations between interparental conflict and blame appraisals. For example, the association between interparental conflict and children's selfblame appraisals was stronger when mothers reported low levels of support for their children's negative emotions; when mothers responded to their children's negative emotions in a dismissive, punitive, or upset way; and when mothers were more coercive. Another pathway by which exposure to interparental conflict might adversely affect adjustment is by undermining adolescents' sense of self-efficacy (Fosco & Feinberg, 2015). Adolescence is an age marked by a striving for a sense of autonomy (Bandura, 1994); it is also an age when individuals are confronted with many challenges (Laursen & Collins, 2009). Self-efficacy is the belief that one is capable of achieving one's goals, despite challenges (Bandura, 1994). It has been described by some as the foundation for well-being (Pajares, 2006). Fosco and Feinberg (2015) found evidence for a developmental cascade model, in which interparental conflict was positively linked to adolescents' threat appraisals, which were negatively related to adolescents' self-efficacy beliefs. In turn, self-efficacy beliefs were associated with adolescent adjustment (subjective well-being, internalizing problems, externalizing problems). Fosco, DeBoard, and Grych (2007) proposed several modifications to the cognitive-contextual framework to better represent children's experiences in families characterized by IPV, especially in families where the violence is frequent and severe. For example, the scope of appraisals was expanded. Specifically, in addition to the threat, blame, and coping appraisals described in the original model, children are hypothesized to appraise interparental conflict (and the specific behavioral tactics, such as violence) along the degree to which it is justified or acceptable. Blame appraisals also include self-blame for failure to protect family members from getting hurt by the violence (e.g., protecting the mother from the father). In addition, some of the modifications put a greater emphasis on how externalizing problems might emerge within the context of IPV. For example, the meaning children ascribe to interparental conflict is theorized to affect not only their behavior but how they regulate their emotions and how they perceive various family members and close relationships within the family. Drawing from the work of Jenkins (2000), children exposed to interparental conflict that is characterized by anger are hypothesized to develop patterns of emotional responding that are organized around

the emotion of anger. That is, anger becomes the default or automatic response to upsetting stimuli, and this begins to occur even in situations when anger is not typically elicited. These children, over time, are hypothesized to develop automatic anger responses to conflicts with others. The modifications to the cognitive contextual framework have not yet received much research attention within families characterized by frequent and severe IPV. However, some of the modifications are consistent with the results of empirical evaluations of tenets of social cognitive theories, indicating that attitudes about the acceptability of violence mediate associations between IPV and both child aggressive behavior and adolescent violence in romantic relationships. Similarly, several studies now indicate that the experience and expression of anger mediates the association between exposure to IPV and adolescent violence in romantic relationships (Kinsfogel & Grych, 2004; Wolfe, Wekerle, Scott, Straatman, & Grasley, 2004). Emotional Security Hypothesis Theory According to Davies and Cummings's (1994) emotional security hypothesis and subsequent writings about emotional security theory (EST) by Cummings and Davies and colleagues, the concept of children's emotional security in the interparental relationship is central to understanding how interparental conflict affects child adjustment. Children are hypothesized to have a higher-order goal of wanting their parents' marital relationship to be stable. That is, in order to preserve a feeling of being safe, secure, and protected within their family, children need to feel confident that their parents can effectively work out their differences and resolve disagreements. Children who are emotionally secure in the interparental relationship have confidence in the stability of their parents' relationship with one another. They have confidence that interparental conflict will not threaten this stability; nor will it adversely affect children's own physical or psychological well-being. Children's emotional security in the interparental relationship is conceptually distinct from their emotional security in the parent-child relationship. Children can be insecure about the interparental relationship but secure in their parent-child relationships, and vice versa. Children's emotional insecurity (or difficulties in preserving a sense of security) in the interparental relationship is inferred from three interrelated component processes, each of which is a type of response to interparental conflict. First, there might be generalized negative emotional reactivity (e.g., fear, anger, and sadness in response to the conflict). As suggested earlier, children often respond to displays of destructive conflict with negative emotions. However, when children are emotionally insecure in the interparental relationship, the negative emotions are likely to be heightened and more persistent. Second, interparental conflict might prompt children to behave in a manner that reduces either their parents' conflict or at least the children's exposure to it. Examples of such behaviors might include antisocial behaviors (yelling at or arguing with a parent), avoidance behaviors to minimize exposure (running away from conflict incidents), and even behaviors that might be construed as prosocial or helpful (trying to help solve the disagreement, trying to comfort one or both parents). Third, interparental conflict may result in internalized representations or working models of

relationship functioning. This might include pessimistic expectations about the implications of the interparental conflict that children may hold for themselves and their families (e.g., the conflict will lead to the breakup of the family). These three component processes have been referred to as the “ABCs” of children's responses to interparental conflict (Cummings & Davies, 2010), referring to affect, behavior, and cognition. It is not necessary for each of the three component processes to be operating to infer emotional insecurity. Rather, these three responses are hypothesized to be distinct but interrelated. These three different responses to interparental conflict may help children address the immediate threats posed by the conflict. However, these same responses may also lead to externalizing or internalizing adjustment problems through a variety of pathways and processes. For example, by definition, negative emotional reactivity is distressing to children, and a prolonged state of such emotions, triggered by prolonged exposure to destructive interparental conflict, or even a cognitive preoccupation with the conflict, can drain children emotionally. As a result, the capacity to draw on emotional resources when dealing with other important developmental challenges is diminished. In addition, certain behavioral responses to interparental conflict might turn into patterns of behavior that generalize to other contexts in which they are not appropriate. For example, Davies and Cummings (1994) made reference to R. E. Emery's (1989) model—on how children who attempt to regulate their parents' conflict may become aggressive or disruptive during marital disputes, and how this behavior might be negatively reinforced. If this family process is repeated, it might result in a more persistent and generalized pattern of child antisocial behavior. Thus, in the long run, children's emotional insecurity in the interparental relationship is expected to confer risk for adjustment difficulties. Consistent with the assertion that children's emotional insecurity in the interparental relationship leads to child adjustment problems, results from many studies link the individual components from which emotional insecurity in the interparental relationship is inferred— emotional reactivity to conflict, regulation of conflict exposure, and internal representations of the marital relationship—with children's externalizing and internalizing problems (e.g., Davies & Cummings, 1998; Davies, Forman, Rasi, & Stevens, 2002; Davis, Hops, Alpert, & Sheeber, 1998; Harold & Conger, 1997). Furthermore, individual components of children's emotional security, as well as latent variables constructed from these individual components, have been found to mediate the relation between interparental conflict and child adjustment (Cummings, Schermerhorn, Davies, Goeke-Morey, & Cummings, 2006; Davies et al., 2002; El-Sheikh, Cummings, Kouros, Elmore-Staton, & Buckhalt, 2008; Kouros, Merrilees, & Cummings, 2008; Mann & Gilliom, 2004). Consistent with EST, different family contextual variables can influence how interparental conflict relates to children's emotional insecurity in the interparental relationship. For example, Kouros, Merrilees, and Cummings (2008) found the link between interparental conflict and children's emotional insecurity to be stronger when fathers reported higher levels of depressive symptoms. Davies and colleagues (2002) showed how several different family variables moderate associations between interparental conflict and emotional insecurity. For example, parenting difficulties and parent-child attachment insecurity led to stronger relations between interparental conflict and children's emotional insecurity in the interparental

relationship. Family cohesiveness, in contrast, led to weaker relations. Another pathway by which children's emotional insecurity in the interparental relationship may lead to child externalizing and internalizing problems is by adversely affecting children's sleep (El-Sheikh, Buckhalt, Cummings, & Keller, 2007). Specifically, children's emotional insecurity in the interparental relationship has been posited to interfere with the quality and duration of children's sleep. When children do not obtain enough high-quality sleep, it negatively influences their day-to-day functioning (Kouros & El-Sheikh, 2015). They are more irritable, have difficulties concentrating, and are not motivated to follow through on tasks and responsibilities, which are all symptoms of different forms of psychopathology (e.g., depressive disorders, oppositional defiant disorder). In families in which there are frequent or prolonged periods of destructive interparental conflict, there may also be prolonged sleep disruptions, contributing to a more general pattern of disrupted behavioral and emotional functioning. Emotional insecurity in the interparental relationship mediates links between destructive interparental conflict and indices of the quality of children's sleep (Kelly & ElSheikh, 2013). In addition, sleep disruptions mediate links between children's emotional insecurity in the interparental relationship and a variety of child adjustment difficulties (ElSheikh et al., 2007). Advances in the Formulation of EST Davies and Sturge-Apple (2007) make use of concepts from ethology and evolution to refine and extend certain aspects of EST, and they articulate more specific hypotheses regarding components of EST. Specifically, in the original formulation of EST, the three indicators of emotional insecurity (generalized negative emotional reactivity, regulation of conflict exposure, and pessimistic expectations about the implications of the interparental conflict for themselves and their families) were conceptualized as linearly related to one another, with each reflecting the common goal of preserving security in the interparental relationship. In the new formulation, the pattern of response across the three indicators is given much more emphasis than it was in the original account of EST. That is, different profiles of children's emotional security/insecurity to interparental conflict are hypothesized (e.g., secure, mobilizing, dominant, caregiving). These profiles are distinguished from one another by the specific emotions, cognitions, and behaviors included in reactions to conflict. They are also distinguished by their origins (a complex function of current ecological conditions, individual developmental histories, and prior experiences with interparental conflict) and by the types of adjustment problems most likely to emerge as a result of these pattern-based responses. For example, the secure profile is characterized by well-regulated, low-level distress reactions to interparental conflict. Children with this pattern are hypothesized to have confidence in their parents' relationship in a way that maintains family harmony. This pattern is theorized to evolve, in part, from witnessing well-managed parental disputes in the context of cohesive interparental and family relationships. These children are not expected to develop significant adjustment problems as a consequence of witnessing destructive bouts of interparental conflict. Children with the dominant profile, in contrast, would react to interparental conflict with high levels of behavioral distress, loss of control, dysregulation, aggression, low levels of

subjective distress, and appraisals of threat. They are theorized to have been exposed to elevated levels of intense, poorly resolved interparental conflict, family disengagement, interparental dissatisfaction, parental dysphoria, avoidant parent-child attachment, and parental physical abuse and neglect. Over time, they would be expected to develop generalized patterns of aggressive, antisocial behavior. There are a number of additional elaborations as well. To offer several examples, the new formulation provides an evolutionary basis for expecting certain types of interparental conflict to be frightening and harmful to children. Specifically, many evolutionary models suggest that humans are predisposed to respond with fear to stimuli that were significant threats to our ancestors. These stimuli may include angry faces, which are much more likely to be observed during conflict characterized by hostile, angry, and aggressive interchanges than by dysphoric, withdrawn, or indifferent interactions. The ability to recognize fear in others is also likely to have been evolutionarily adaptive, because others' fear responses likely signaled impending danger. Accordingly, observing one parent being fearful of another may be particularly likely to pose a threat to children's emotional security and, over time, their adjustment. The new formulation also emphasizes the primacy of initial, nonconscious, automatic processing of threat cues and affective responses to those cues, followed over time by the development of responses that operate under some cognitive control. Thus, this aspect of the reformulation implies that fearful affect plays a primary role as a cause of cognition, which distinguishes this theory from those positing that children's cognitive appraisals of a particular conflict incident mediate the link between that incident and their emotions. In addition, the new formulation downplays the role of constructive interparental conflict in contributing to children's emotional security. There is already some empirical support for aspects of this new formulation of EST. Indeed, findings by Davies and Forman (2002) on different profiles in children's immediate responses to interparental conflict helped prompt the reformulation. More recently, Davies and colleagues (2012) found that when destructive and constructive conflict behaviors are considered simultaneously, destructive conflict is associated with children's emotional insecurity but constructive conflict is not, even though it does contribute to future child adjustment. Specific Emotions Theory Similar to the cognitive-contextual framework and EST, specific emotions theory (Crockenberg & Langrock, 2001) also holds that the effect of interparental conflict on children depends on children's interpretation of what the conflict means for themselves and their family. However, the focus is on the children's interpretation of how the conflict affects their ability to attain certain goals. Children's goals vis-à-vis interparental conflict can be broad, such as attaining security in the interparental relationship (as in EST), or specific, such as being able to invite a friend over after school. The meaning the conflict has for goal attainment leads children to experience specific emotional reactions to the conflict. For example, when the conflict is perceived to block a goal or make it unattainable, children are theorized to respond with anger. Repeated blocking of goals because of interparental conflict is expected to lead to

sadness as well. When the interparental conflict poses a looming but uncertain threat to the children's goals, fear is the consequence. Children can, of course, have multiple goals simultaneously, and they may experience multiple emotions in reaction to any given conflict (e.g., fear over insecurity in the interparental relationship and anger because plans for a family activity are canceled because of the conflict). Over time, the patterns of children's reactions to conflict are thought to be reflected in more generalized patterns of adjustment. Thus, angry reactions translate to a broader pattern of aggressive or externalizing behavior problems, whereas fearful and sad reactions translate to internalizing problems. Recognizing that multiple reactions can occur simultaneously, specific emotions theory posits that the co-occurrence of different emotions influences how each is expressed. For example, when angry reactions to interparental conflict are accompanied by fear, the fear may reduce the likelihood of aggression or increase the likelihood of withdrawal or avoidance, thus setting the stage for the development of internalizing symptoms. Specific emotions theory hypothesizes a second direct-effects pathway, which involves the same-sex modeling process derived from social cognitive theory. A same-sex modeling hypothesis specifies that children disproportionately attend to and display the behavior of their same-sex parent, and by the age of 5 or 6 years, children have developed gendered scripts that have been incorporated into their personal identities (Levy & Fivush, 1993). Thus, during incidents of interparental conflict, boys are expected to be more likely to attend to their fathers than to their mothers. The opposite is expected of girls. Similarly, to the extent that fathers' and mothers' behavior differs during incidents of conflict, boys' behavior should be more parallel to that of their fathers, whereas girls' behavior should be more parallel to that of their mothers. Data testing this theory are sparse. However, there is empirical evidence consistent with some of its predictions. For example, specific emotions experienced by boys in response to interparental conflict have been found to interact with the fathers' behavior to predict specific types of adjustment problems (Crockenberg & Langrock, 2001). Specifically, boys' anger in response to interparental conflict interacted with fathers' maritally aggressive behaviors to predict boys' externalizing problems. Similarly, boys' fear in response to interparental conflict interacted with fathers' maritally aggressive behaviors to predict boys' internalizing problems. In contrast, girls' specific emotions in response to interparental conflict did not interact with fathers' martially aggressive behaviors to predict girls' adjustment problems. In addition, boys' emotions in response to interparental conflict did not interact with mothers' maritally aggressive behaviors to predict boys' adjustment problems. Interestingly, for girls, the associations predicted by this theory failed to emerge, raising questions about its applicability for both boys and girls. Trauma Theory Individuals can experience traumatic stress symptoms after experiencing events that pose a significant threat of physical injury, severe harm, or death to an individual or others, and which evoke intense fear. In some circumstances, interparental conflict can be considered a traumatic stressor. Considerable evidence indicates that children exposed to IPV—which by many definitions poses a threat of physical harm to a child's caregiver and thus is frightening to a

child—are at elevated risk for developing symptoms of posttraumatic stress (Evans, Davies, & DiLillo, 2008; Graham-Bermann & Levendosky, 1998; Kitzmann, Gaylord, Holt, & Kenny, 2003). Moreover, in many families marked by IPV, the family violence occurs repeatedly. Sustained or repeated exposure to multiple traumatic events, known as complex trauma, is associated with particularly detrimental child outcomes (Cloitre et al., 2009; Margolin & Vickerman, 2007). A large body of research has outlined ways in which the development and functioning of biological stress-response systems may be disrupted by exposure to significant and/or prolonged stress (De Bellis, 1999; Gunnar & Quevedo, 2007). The hypothalamic-pituitaryadrenal (HPA) axis is a neuroendocrine system that helps govern the organism's response to stress. Exposure to IPV can lead to dysregulated HPA axis functioning, including alterations in cortisol reactivity and diurnal patterns of cortisol release (Davies, Sturge-Apple, Cicchetti, Manning, & Zale, 2009; Saltzman, Holden, & Holahan, 2005; Davies, Sturge-Apple, Cicchetti, & Cummings, 2007). Increased cortisol can result in neuronal damage and in other changes in HPA functioning and secretion of neurohormones that can prompt pruning of receptors for those hormones, thus physically altering the functioning of the entire system. Such alterations compromise the child's physiological responses to stress going forward. The sympathetic nervous system (SNS) is also activated by traumatic events, resulting in increases in the level of circulating catecholamines (e.g., norepinephrine). Such increases not only increase arousal (i.e., the fight-or-flight syndrome) but over time can also result in pruning of receptors that would dampen their effect. Over time, such lasting physical changes in the structure and functioning of the stress-response systems, especially while those systems are still developing, hold serious implications for a child's future abilities to respond adaptively to stress. There is substantial theory on the roles of particular psychological and social factors in the relation between exposure to a significant stressor and the likelihood of traumatic stress symptoms. Independent of the IPV itself, each child's unique configuration of characteristics shapes how he or she responds to it; these characteristics include the child's genetics, temperament, attachment functioning, existing psychopathology, functioning of the stressresponse system, and developmental competencies in multiple domains (i.e., affective, cognitive, behavioral, and social). These both influence and, over time, are influenced by the child's efforts to respond and adapt to the stress and distress that exposure to severe forms of interparental conflict, such as IPV, entails. Meta-analytic findings confirm theoretical propositions that factors subsequent to an episode of IPV exposure also figure heavily in a child's risk of developing posttraumatic stress symptoms (Trickey, Siddaway, Meiser-Stedman, Serpell, & Field, 2012). These include factors specific to the IPV, such as everyday reminders or harbingers of it, its reoccurrence and patterning of occurrence, and the extent to which it may lead to or be associated with other potentially traumatic stressors, such as child maltreatment. They also include the child's processing of the event and its consequences and the child's attributions about his or her own emotions, behavior, and coping during and after the event. In addition, because IPV tends to recur, living under an ever-present threat that violence may erupt may itself be a traumatic stressor (Margolin & Vickerman, 2007).

Indirect-Effects Models Parenting Consistent with decades of research on child development and developmental psychopathology, most of the direct-effect theories just reviewed acknowledge that the quality of the parent-child relationship is an important determinant of child adjustment. In these theories, aspects of parenting or the parent-child relationship are presumed to partially mediate the link between interparental conflict and child adjustment problems. Similarly, most family systems theories consider the association between interparental conflict and child adjustment to be a function of disrupted family processes, including parenting and the parent-child relationship. In this section of the chapter, we briefly review some of the empirical evidence on the relation of interparental conflict to parenting and to parent-child relationship variables. We then describe theory explaining the relation. For simplicity, we use the term parenting to refer to the multitude of characteristics, behaviors, and processes considered in this research. Results of numerous studies show interparental conflict to be associated with both broad dimensions of parenting (e.g., acceptance/rejection) as well as with specific parenting behaviors, ranging from the amount of general conversation that parents engage in with their children to harsh and physically abusive parenting behaviors. In addition, changes in parenting that are associated with interparental conflict can occur quickly, even during or shortly after a conflict incident (seconds, minutes, days). Or they can unfold slowly over time, becoming fully manifest only months later. Experimental and well-controlled longitudinal studies provide evidence of causation: that interparental conflict causes alterations in parenting. For example, Jouriles and Farris (1992) randomly assigned parents to discuss either a highly conflictual or nonconflictual topic for them as a couple. During the discussion, their child was in another room, playing a game that involved making a substantial mess. Immediately after the couples' discussion, one (randomly chosen) parent, who had been instructed to have the child clean up the mess, joined the child in the playroom. Results indicated that the nature of the parents' discussion with one another influenced parent-child interaction during the cleanup task. Specifically, the level of general conversation (parent-child chitchat about matters unrelated to cleaning up the room) was consistently high after the nonconflictual couple interaction. In contrast, it was noticeably lower immediately following the conflictual couple interaction, gradually increasing over the course of the parent-child interaction. In addition, fathers gave more vague and confusing instructions to the child immediately following the conflictual couple interaction. There is now also a substantial body of research suggesting that parenting mediates the relation between interparental conflict and child adjustment problems. In fact, the results of several studies suggest that when multiple dimensions of parenting are considered, parenting sometimes fully mediates the relation. For example, Fauber, Forehand, Thomas, and Wierson (1990) used data from parents and adolescents to form latent variables of parental psychological control, rejection/withdrawal, and lax control. When these were considered simultaneously, psychological control and rejection/withdrawal fully mediated the link between interparental conflict and child internalizing problems and partially mediated the link

between interparental conflict and child externalizing problems. Harold, Fincham, Osborne, and Conger (1997) used children's reports across several measures of parent-child interaction to construct constructs reflecting mother-child and father-child hostility. These constructs fully mediated the link between interparental conflict and child externalizing problems and partially mediated the link between interparental conflict and child internalizing symptoms. Similarly, Gonzales, Pitts, Hill, and Roosa (2000) collected children's reports of parental acceptance and support (similar to the rejection/withdrawal variable of Fauber et al., 1990), inconsistent discipline, and hostile control (similar to the psychological control variable of Fauber et al., 1990, and the hostility measures of Harold et al., 1997). When these parenting variables were considered simultaneously in a mediation model, parental acceptance and support and parental inconsistent discipline fully mediated the link between child reports of interparental conflict and child reports of child adjustment problems. However, the mediation effects were not replicated when using parent reports of child adjustment problems. A number of explanations have been offered for the link between interparental conflict and parenting. The spillover hypothesis, or some variant thereof, is perhaps the most widely acknowledged process for explaining how interparental conflict affects parenting (see Erel & Burman, 1995; Krishnakumar & Buehler, 2000, for reviews). According to this hypothesis, negative emotions and behavior in the interparental relationship permeate or spill over into the parent-child relationship. For example, overt anger or hostility between parents during a conflictual interaction may spill over into the parent-child relationship, leading to anger and hostility toward the child, including overly harsh and physically aggressive discipline (Appel & Holden, 1998; Jouriles, McDonald, Slep, Heyman, & Garrido, 2008). Patterson (1982) described how interparental conflict alters parents' moods and, consequently, disrupts many parenting behaviors that could contribute to the development of child adjustment problems. These include monitoring children's activities and whereabouts, consistent enforcement of rules, delivery of contingent consequences for child misbehavior, and effective problem solving. Similarly, frustration induced during interparental conflict may contribute to frustration in the parent-child relationship, which manifests itself as a low tolerance for child misbehavior and the use of power-assertive child management techniques (Gerard, Krishnakumar, & Buehler, 2006). It has also been hypothesized that the preoccupation and emotional fatigue generated by interparental conflict influences the degree to which parents can be warm, supportive, and emotionally available for their children (e.g., Fosco et al., 2007). The spillover hypothesis is often conceptualized in the context of a micromodel of family interaction (e.g., Jouriles & Farris, 1992; Patterson, 1982). That is, an episode of interparental conflict generates negative affect in the parents, and these emotions influence how the parents interact with their children either shortly after the conflict episode (minutes, hours, or days), or sometimes even during the conflict itself (e.g., when a child intervenes to attempt to stop the conflict). The spillover hypothesis, however, has also been conceptualized to play out over longer periods of time. For example, a variant of the spillover hypothesis is that interparental conflict contributes over time to heightened levels of parenting stress or diminished parental mental health, which then interferes with the parents' capacity to parent optimally (Levendosky & Graham-Bermann, 2001; Levendosky, Huth-Bocks, Shapiro, & Semel, 2003). Consistent

with this hypothesis, mothers' mental health symptoms have been found to mediate associations between mothers' experience of IPV during pregnancy and preschoolers' adjustments problems (Flach et al., 2011; Levendosky et al., 2006). Investigators have also used the spillover hypothesis, or some variation of it, to link interparental conflict to the erosion of the quality of the parent-child relationship (Davies & Sturge-Apple, 2007; Owens & Cox, 1997). For example, as suggested by the spillover hypothesis, the residual distraction and distress from interparental conflict detracts from parents' emotional availability to their children. Over time, the parents' emotional unavailability may diminish the child's sense of the parents as reliable sources of support and comfort. In addition, witnessing a parent's distress, fear, anger, or loss of emotional control may also alter the child's perception of the parent as a source of safety, strength, or protection. In contrast to the spillover hypothesis, the compensatory hypothesis posits that parents attempt to compensate for negativity in the interparental relationship by investing more heavily in the parent-child relationship. Nevertheless, heavy investment in the parent-child relationship is not always beneficial for children. For example, detrimental effects may emerge if parents view the parent-child relationship as a substitute for the interparental relationship. In such cases, children may be burdened by feelings that they must emotionally support their parents as adult partners typically support one another (Peris, Goeke-Morey, Cummings, & Emery, 2008). Some studies using observational measures of parenting, however, have found associations between interparental conflict and more empathic parenting (Belsky, Youngblade, Rovine, & Volling, 1991; Mahoney, Boggio, & Jouriles, 1996). Similarly, in one qualitative study, 20% of battered women reported that their experiences with violence had some positive effects on their parenting, such as giving them greater empathy for their children (Levendosky, Lynch, & Graham-Bermann, 2000). Although the spillover and compensatory hypotheses are both employed as explanations of the link between interparental conflict and parenting, testing these hypotheses can be very challenging, partly because the optimal time frame for capturing these processes is not clear from theory or research. In fact, it might be argued that both processes can operate simultaneously within the same family, or even the same parent, at different points in time. In a recent test of the spillover and compensatory hypotheses, Kouros, Papp, Goeke-Morey, and Cummings (2014) had parents independently rate the emotional quality of their relationship with their spouse and with their child each day for 15 days. The authors found that fathers' daily ratings of poor marital quality predicted lower father-child relationship quality on the same day, which is consistent with the spillover hypothesis. In contrast, mothers showed improvement in the quality of the mother-child relationship when marital quality was rated poorly on the previous day, which is consistent with the compensatory hypothesis. Weakening of Boundaries between Family Subsystems According to family systems theory, clear boundaries between the parents' relationship with one another (the marital subsystem) and the parents' relationships with their children (the parent-child subsystems) are crucial to healthy family functioning (Kerig, 1995; Minuchin,

1974). Clear boundaries establish roles for family members that are developmentally and socially appropriate and are sensitive and appropriately responsive to the emotional needs of family members. In the absence of clear boundaries, children may be cast in roles that are inappropriate for the developmental level (e.g., the parentified child) and that do not meet their emotional needs. Family systems theorists suggest that interparental conflict weakens boundaries between the marital and parent-child subsystems and, as a consequence, influences the development of child adjustment problems. Weak boundaries between the marital and parent-child subsystems might be conceptualized as a type of parenting deficit or problem. That is, it might be argued that it is the parents' responsibility to establish and maintain clear boundaries between these two subsystems. Parents blur the boundaries when they draw the children into their interparental conflicts. Examples of this include when one parent engages in behaviors to attempt to recruit children into an alliance or coalition against the other parent. These processes have been given a number of labels in the family systems literature, including triangulation and the formation of cross-generational alliances. Blurring the boundaries between marital and parent-child subsystems might involve circumstances in which one parent talks badly about the other parent to the children or offers special privileges and gifts in an effort to get the child to take the side of a parent. Boundaries are also blurred when parents let their children intervene or involve themselves into their conflicts. Children might try to do this for a variety of reasons. For example, they might blame themselves for their parents' conflicts (especially when the conflict is about them or something they did) and thus feel responsible for stopping such conflicts. Another example is that in violent families, children may feel a need to protect one or both parents from physical harm. Thus, they intervene to prevent the conflict from escalating to a point where a parent becomes injured. Parents might also blur the boundaries between these two family subsystems as part of the aftermath of interparental conflict. Examples include when they turn to their children to provide them with support and comfort; such a process has been labeled as parentification, adultification, and role reversal. Support and comfort might be sought to deal with the troubled marriage or for addressing other life stressors, because the parent does not receive support and comfort from his or her spouse. There is evidence from prospective research that the blurring of boundaries between the marital and parent-child subsystems contribute to child problems, as well as processes that are likely to lead to child problems. For example, Fosco and Grych (2010) found triangulation related positively to parent-adolescent conflict over time. Buehler, Franck, and Cook (2009) found positive associations between triangulation and adolescent externalizing and internalizing problems, as well as peer rejection. Coparenting Another indirect-effects model, which overlaps with parenting and family systems models, emphasizes the construct of coparenting and the extent to which mothers and fathers cooperate and support each other in raising their children. McHale and colleagues (McHale, Kuersten, & Lauretti, 1996) propose two types of coparenting dynamics, overt and covert, both of which can be affected by interparental conflict. Overt coparenting dynamics involve interactions

between the parents about the child. Interparental conflict may result in one parent actively interfering with or undermining the other parent's interactions with the child, which may leave the child confused or compelled to defend or choose one parent over the other. Alternatively, when coparenting is disrupted by interparental conflict, one caregiver may withdraw from, reduce the frequency of, or otherwise alter the nature of interactions with the child (McHale et al., 1996; Katz & Gottman, 1996). In contrast, covert coparenting occurs when a parent delivers messages about the other parent while alone with the child (McHale & Ramussen, 1998). Ideally, such messages are positive and reflect a strong parental alliance. Coparenting that has been corrupted by interparental conflict can be seen in interactions in which one parent casts the other in a negative light, attempts to persuade the child to take sides against the other parent, or attempts to win the child's affection at the other parent's expense. Although research is sparse, there is some evidence that overt coparenting processes mediate the link between interparental conflict and child adjustment problems. For example, Katz and Low (2004) found that hostile-withdrawn coparenting behaviors mediate associations between IPV and children's anxiety and depression. Stroud, Meyers, Wilson, and Durbin (2015) found marital functioning to be related to children's internalizing and externalizing problems through reductions in coparenting warmth. Other Explanations and Considerations Cox, Paley, and Harter (2001) theorize that through various family and parenting processes, interparental conflict may strain sibling relationships, which contributes to the development of child adjustment problems. For example, elaborating on the family systems theory concepts of triangulation and the formation of cross-generational alliances, it seems plausible that if a child is aligned with one parent against the other, that child is likely to receive preferential treatment from the parent with whom the alliance is formed. Preferential treatment of children can strain sibling relationships and influence how one sibling treats the other. Because sibling interactions provide children opportunities to develop social interaction (e.g., conflict management) and coping skills (E. R. Anderson, Hetherington, Reiss, & Howe, 1994), strained relationships can influence the development of these skills and thus the development of adjustment problems. Alternatively, interparental conflict may lead to behavior problems in one child, which then contribute to the development of behavior problems of one or more siblings. For example, older siblings typically exert more influence on younger siblings than vice versa (Stoneman & Brody, 1993). If interparental conflict leads to adjustment problems in an older child, that child's problems might then adversely affect younger siblings (Erel, Margolin, & John, 1998). Research on interparental conflict and sibling relationships is sparse, but there is some evidence consistent with the idea that sibling relationships may be an important variable in the link between interparental conflict and child adjustment (Piotrowski, Tailor, & Cormier, 2014; Tailor, Stewart-Tufescu, & Piotrowski, 2015). Noting the high degree of co-occurrence for violence between intimate adult partners and violence directed at children, Slep and O'Leary (2001) proposed an integrated theoretical framework for family violence, drawing from the largely separate research literatures on risk factors for IPV and risk factors for physical child abuse. Slep and O'Leary highlighted the fact

that there is considerable overlap in the risk factors for the perpetration of each of these forms of family violence, including individual-level variables (e.g., growing up with violence in their own family of origin, problems with impulsivity, problems with substance use and abuse, depressive symptoms) as well as an assortment of family- and community-level variables (e.g., economic strain, stressful events, social isolation). The collective presence of these variables is hypothesized to lay the groundwork for an individual to direct violence toward multiple family members—anyone who happens to provoke the high-risk individual. Tests of this integrated theoretical framework indicate that individuals who engage in both IPV and violence directed at children, compared to individuals who engage in only one of these forms of family violence or no violence at all, report higher collective levels of these risk factors (Slep & O'Leary, 2009). Interparental conflict and IPV have also been associated with dimensions of peer interaction and peer relations. To offer a couple of examples, investigators have found interparental conflict to be positively associated with friendship instability and indirectly related to friendship quality (Schwarz, Stutz, & Ledermann, 2012). IPV has been found to be positively associated with peer victimization and peer bullying (Knous-Westfall, Ehrensaft, MacDonell, & Cohen, 2012). Additional research is necessary to evaluate whether peer relations mediate links between interparental conflict and child adjustment problems.

Child Attributes and Family/Community Factors Contributing to Individual Differences There is considerable variability in children's adjustment within families marked by interparental conflict. For example, even in samples of families characterized by extreme levels of conflict, such as IPV that is frequent and severe enough to prompt a parent (usually the mother) to seek refuge at a domestic violence shelter, there are marked differences in child adjustment. Some of the children develop externalizing problems, others develop internalizing problems or both externalizing and internalizing problems, and still others develop few or no problems (Grych, Jouriles, Swank, McDonald, & Norwood, 2000). This is true not only across children in different families but even among siblings within the same family (Skopp, McDonald, Manke, & Jouriles, 2005). A developmental psychopathology perspective conceptualizes individual development as the interplay between a changing organism growing up in a changing context (Cicchetti, 1993; Davies & Cicchetti, 2004). This conceptualization implies that the development of child adjustment problems within the context of interparental conflict varies as a function of changes occurring within the child (e.g., maturation), within the interparental relationship, and in the child's family and social environment. In this section of the chapter, we describe how individual attributes of the child might influence the way that interparental conflict adversely affects child adjustment. These attributes include child sex, developmental considerations, immediate behavioral responses to interparental conflict, characteristic coping behaviors and cognitive styles, neurobiological responses to interparental conflict, and genetic influences. We

then consider aspects of the broader family and community context that might influence the effect of interparental conflict on child adjustment.

Child Sex Child sex has received quite a bit of attention in this research literature, in part because early research on the association between marital and child problems suggested that the association was stronger for boys than for girls (Grych & Fincham, 1990; Reid & Crisafulli, 1990). However, continued research, particularly with relatively large and representative community samples (e.g., Jouriles, Bourg, & Farris, 1991), called this conclusion into question, and later meta-analytic reviews failed to support the notion that child sex moderates the association between interparental conflict and child adjustment problem (Buehler, Anthony, Krishnakumar, & Stone, 1997). In addition, child sex also does not appear to moderate the association between many hypothesized mediators of the link between interparental conflict (e.g., threat and self-blame appraisals, negative affect) and child adjustment problems (Rhoades, 2008). This is not to say that child sex is unimportant for understanding the association between interparental conflict and child adjustment. Rather, it is likely that the role played by child sex is complex, and a clear and consistent pattern of findings has not yet emerged. It is our view that other child and family attributes, some of which may be associated with child sex, are likely to be more promising in understanding the heterogeneity in the association between interparental conflict and child adjustment problems.

Developmental Considerations A child's developmental level is often acknowledged as important for understanding the link between interparental conflict and child adjustment. Studies thus far, however, have not evaluated how developmental processes influence this link, beyond examining age as a moderator of specific associations. Changes occur in children's competencies across multiple domains (cognitive, emotional, social, physical) in the course of normal development, and these changes are relevant to how children are affected by interparental conflict. To illustrate, next we briefly discuss how normal developmental changes in cognition might be salient to concepts from theories on the link between interparental conflict and child adjustment. We also discuss developmental transition periods that might be important and conclude with a discussion on why greater attention to developmental processes is needed to guide the next generation of theory and research on interparental conflict and child adjustment. According to Piaget's theory of cognitive development, children rely on their sensory and perceptual abilities to explore their environment and make sense of their world. Infants quickly begin to form schemas—internal representations of people, objects, and events—to organize their experiences. These schemas may be very simple at first but become increasingly sophisticated as the child assimilates new experiences into existing schemas and alters existing schemas to accommodate new experiences. A schema of one's caregiver is one of the earliest to develop; based on their experiences, infants develop an internal working model of their caregiver that allows them to predict how responsive, reliable, and trustworthy she or he is and will be, especially during times of stress.

Several of the direct-effect models reviewed earlier point to the importance of children's schemas about their parents' relationship with one another. For example, both the cognitivecontextual framework and EST posit that children create internal representations of the interparental relationship, which are based on their experiences within the family, including their exposure to interparental conflict. Both theories also hold that children become distressed when they perceive the security or stability of the interparental relationship to be threatened. When sustained or repeated over time, this distress can lead to more pervasive adjustment problems. Based on Piaget's theory on children's cognitive development, it is reasonable to assume that internal representations of the interparental relationship begin to form early in a child's life, but the definition of early—and, by implication, late—is unclear. It is also not clear when key aspects of these internal representations are likely to develop, such as when children begin to view their parents' relationship as something that is supposed to be stable, and when this view itself crystallizes, becoming unwavering over time. It might be argued that this information is key to understanding the ages or developmental stages at which EST and the cognitive-contextual framework become applicable for understanding the association between interparental conflict and child adjustment. According to the cognitive-contextual framework, children's appraisals about interparental conflict (threat and self-blame appraisals) are key to understanding how the conflict affects children. However, young children lack the cognitive abilities necessary for making sophisticated appraisals and attributions. Before age 7 or so, children typically do not understand that others do not see the world from their vantage point. They may be able to reason deductively but can apply those reasoning skills only to real, concrete scenarios. Thus, they use only concrete characteristics of events to make attributions. At around age 11, when children enter the formal operational stage of cognitive development, they begin to learn to think abstractly and reason about hypothetical problems. It is only then that they become able to form sophisticated attributions and appraisals. Some have argued that appraisals of interparental conflict are more salient predictors of children's adjustment when children are 10 years of age or older, as compared to younger age groups. For example, in a sample of children who were living with extreme IPV, Jouriles, Spiller, Stephens, McDonald, and Swank (2000) found that younger children (ages 8–9) made more self-blame attributions than older children (ages 10–12), but the relation of self-blame with externalizing and internalizing problems was stronger for the older age group. In a metaanalytic review, Rhoades (2008) found that children's threat and self-blame appraisals were more strongly associated with their externalizing and internalizing problems after the age of 10. This is not to say, though, that younger children's appraisals of interparental conflict are unimportant or not useful. McDonald and Grych (2006), for example, found that children's selfblame appraisals accounted for the relation between interparental conflict and child internalizing problems in a sample of 7- to 9-year-olds. Ablow, Measelle, Cowan, and Cowan (2009) found that 5-year-olds' self-blame appraisals mediated relations between interparental conflict and child internalizing problems. Taken together, this body of research suggests that children's appraisals of interparental conflict are salient predictors of their adjustment problems, but that more robust and stronger associations are likely to be found in samples of

children over the age of about 10. It is also important to consider normal developmental transition periods, such as the transition to school or the onset of puberty. Jaffee and Poulton (2006) proposed that the relation between mothers' depression and child adjustment should be strongest during transition periods because children are more reliant on their mothers during those periods to help guide them in adapting to a changing situation, and depressed mothers are less able to offer that guidance. This reasoning can be applied to interparental conflict; interparental conflict that drains parents' emotional resources, increases their negative affect, or otherwise diminishes their ability to offer support (as suggested in our review of indirect-effect models) should be most problematic for child development during periods when children most need parental support to help navigate important transitions. Greater consideration of developmental processes will be important for advancing both theory and research on the topic of interparental conflict and child adjustment. With regard to theory, many of the hypothesized mechanisms linking interparental conflict with child adjustment (e.g., cognitive and emotional abilities) change as a function of development and can differ dramatically across children of different ages. In fact, some of the mechanisms invoked by certain theories likely apply only to children during specific developmental periods. In addition, the way we conceptualize and measure these hypothesized mechanisms needs to be developmentally appropriate. That is, although the underlying mechanisms posited to be important may be the same across development, the manifestation of these mechanisms may change over time. Therefore, the measures we use need to be developmentally sensitive to account for this heterotypic continuity in children's development.

Behavioral Responses to Interparental Conflict (Stress) Several theories of the relation between interparental conflict and child adjustment suggest that children's immediate behavioral responses to conflict are key to understanding how interparental conflict eventually influences child adjustment. More precisely, many of the theories hypothesize that children respond to destructive interparental conflict with distressing emotions or cognitions, which prompt certain behavioral responses. Although there is a wide range of potential behavioral responses, they are often categorized as either involvement or avoidant behaviors (Laumakis, Margolin, & John, 1998; O'Brien, Bahadur, Gee, Balto, & Erber, 1997; Rhoades, 2008; Shelton & Harold, 2008). Involvement behaviors include certain types of antisocial behaviors, such as yelling at the parents or creating a distraction (e.g., fighting with a sibling) to try to get their parents to stop arguing. However, they may also include behaviors that are prosocial in nature, such as attempting to problem-solve with parents or comfort their parents. Common avoidant behaviors include moving to another room or going outside when a conflict erupts. Avoidance has sometimes also been conceptualized to include cognitive avoidance strategies, such as trying not to think about the conflict. Involving themselves in their parents' conflicts puts children in the midst of what may sometimes be emotionally intense, hostile parental interactions. Contagion theories and related data suggest that children exposed to such situations may take on their parents' angry,

aggressive feelings and behaviors themselves (e.g., Cummings & Cummings, 1988; Cummings, Goeke-Morey, & Papp, 2004; Cummings, Iannotti, & Zahn-Waxler, 1985). In addition, some children yell at or argue with a parent to attempt to stop the conflict (Shelton & Harold, 2008), and this aggressive behavior may sometimes succeed in stopping the conflict (Jenkins, Smith, & Graham, 1989). As reviewed earlier under social cognitive theories, this family process in which a child responds to parents' conflict with aggressive behavior and the parents' conflict ceases may reinforce the child's use of aggression, increasing the probability that it will recur and perhaps develop into more generalized aggressive response patterns (Davis, Hops, Alpert, & Sheeber, 1998; R. E. Emery & Kitzmann, 1995; Patterson, 1982). Children's involvement in interparental conflict might also lead to or reflect problematic family processes that can have negative consequences for child development. As reviewed earlier, family systems theories suggest that clear boundaries between the marital and child subsystems define appropriate family roles (e.g., who is the parent and who is the child) and predict healthy child functioning (Kerig, 2005; Minuchin, 1974). Children's attempts to get their parents to stop fighting might be conceptualized as a boundary disturbance or breakdown in the family role structure—a form of role reversal. From this perspective, when children attempt to manage their parents' conflicts, they are assuming an adult or parent role within the family. Doing this burdens children with responsibilities well beyond their years, which can be very stressful and lead to negative outcomes (Johnston, Gonzalez, & Campbell, 1987; Kerig, 2005; Peris et al., 2008). In contrast to involvement in interparental conflict, avoidance behaviors have been theorized to have both positive and negative effects on child adjustment (O'Brien et al., 1995; Shelton & Harold, 2008). For example, successful avoidance shields children from the hostility and coercive and violent tactics sometimes expressed during conflicts or from becoming a target of parental hostility or violence themselves. Because it might be easier for children to modulate their own emotions and thoughts when they physically distance themselves from the conflict, avoidance behaviors may also help children cope with it. Avoidance might also reflect healthy or positive family characteristics. As noted, family systems theorists hypothesize that clear boundaries between the marital and child subsystems predict healthy child functioning (Kerig, 2005; Minuchin, 1974). Avoidance of interparental conflict might thus reflect such clear boundaries. Children's avoidance of interparental conflict, however, might also lead to negative child outcomes through several different processes. Avoidance of interparental conflict might generalize to other potentially distressing events or interactions. Learning theories suggest that avoidance of interparental conflict might be reinforced if it reduces distressing cognitions and emotions prompted by the conflict. Over time, the relief that comes from avoiding exposure to interparental conflict may lead the child to begin to avoid other life events with the potential to cause distress (Rhoades, 2008). The broader research literature on children's stress and coping is consistent with this idea, suggesting that a disengaged coping pattern, one that includes avoidance, increases risk for internalizing and externalizing child problems (Compas, ConnorSmith, Saltzman, Thomsen, & Wadsworth, 2001).

Avoidance of interparental conflict might also interfere with the development of adaptive skills for managing emotions and cognitions under stress. This hypothesis is consistent with challenge models in the developmental psychopathology literature (Davies & Sturge-Apple, 2007), in which children's exposure to stress, in small doses, is hypothesized to have “steeling” effects, especially if these exposures occur in otherwise supportive contexts. Specifically, exposure to a stressor might inoculate children to subsequent exposures by prompting them to develop effective ways of coping with stress and adversity (Garmezy, Masten, & Tellegen, 1984; Rutter, 1987). As applied to interparental conflict, challenge models suggest that exposure to destructive interparental conflict can help children develop effective skills for dealing with it, as well as for dealing with other difficult, stressful situations. Therefore, avoiding interparental conflict might interfere with the development of those skills. Basic developmental processes might play a role in predicting children's behavioral responses to interparental conflict. For example, research on the developmental neurobiology of traumatic stress (Pynoos, Steinberg, Ornitz, & Goenjian, 1997) suggests that it is not until approximately age 8 that children begin to actively entertain thoughts of personally addressing or disarming the source of the stress, in contrast to avoiding or escaping it. Similarly, as their cognitive processes develop, children also begin to weigh the benefits and risks of behaviorally responding to their parents' conflict. In mild conflict, the risk to the child of misbehaving or otherwise intervening to derail or stop the conflict may be low, and this strategy therefore may be effective. The risk of intervening in conflict that involves physical aggression is higher; however, children may nonetheless intervene if they perceive the risk of not intervening to be even greater. Although young children sometimes intervene in order to stop their parents' fighting, the use of this strategy is expected to be more common among older children (preadolescent or adolescent) (Davies, Myers, & Cummings, 1996; Shelton & Harold, 2007). It is also noteworthy that intervening in interparental conflict is more closely associated with child adjustment problems among older (over 10 years), as compared to younger, children (Rhoades, 2008). The empirical literature linking involvement in and avoidance of interparental conflict to child adjustment problems is mixed. Most studies have found involvement and avoidance both to be positively associated with child adjustment problems (e.g., Davies, Forman, et al., 2002; Nicolotti, El-Sheikh, & Whitson, 2003; Shelton & Harold, 2008). However, the opposite has been found, as well, particularly with child avoidance. For example, O'Brien, Margolin, and John (1995) found children's avoidance of interparental conflict to be negatively associated with child anxiety symptoms, and Kerig and colleagues (1998) found girls' avoidance of interparental conflict to be negatively associated with externalizing problems.

Characteristic Coping Behaviors and Cognitive Styles Related to the section on children's immediate behavioral responses to interparental conflict, some children may have characteristic ways of coping with stressful events, which might moderate associations between interparental conflict and child adjustment problems. In fact, some theorists have conceptualized children's immediate behavioral responses to interparental

conflict as coping behaviors (e.g., Kerig, 2001). The few studies that have examined children's coping behaviors as potential moderators of the association between interparental conflict and child adjustment problems have yielded intriguing results. Nicolotti, El-Sheikh, and Whitson (2003), for example, found that a combination of support and active coping (e.g., having other people listen to their feelings about problems; planning or thinking about ways to solve problems) moderated associations between IPV and child adjustment problems. Interestingly, this form of coping served a protective function for girls; higher levels of support and active coping mitigated the association of IPV with depression and self-esteem difficulties. However, this coping style increased the magnitude of these associations for boys. Rogers and Holmbeck (1997) also found the association between IPV and child adjustment problems to be weaker among children who made greater use of social supports. (Sex differences were not examined.) Other characteristic coping behaviors have also been found to moderate associations between interparental conflict and child adjustment problems. Nicolotti and colleagues (2003) found that greater avoidance coping (e.g., avoiding the problem by physically leaving, staying away from it, or imagining it to be better) is associated with a stronger relation between parental IPV and children's externalizing and internalizing problems and self-esteem difficulties. In contrast, distraction coping (e.g., engaging in physical activity or entertainment activities to avoid thinking about the problem) functioned as a protective variable, weakening the association between IPV and child depressive symptoms. Drawing from the broader literature on child coping may help refine our understanding of how children's responses to interparental conflict can contribute to the development of adjustment problems. For example, children's perceptions of how much control they have over their parents' relationship, and their ability to predict interparental conflict (Compas, 2009)—key factors in understanding why psychological stress is harmful to one's mental health—may be important factors to consider as well. The broader literature on psychopathology, particularly on depression, suggests that certain cognitive styles or predispositions interact with the occurrence of stressful environmental events to produce depressive symptomatology. For example, models of depression suggest that a negative attributional style (a predisposition to explain stressful or negative events in a particular way) increases risk for depressive symptoms when negative or stressful life events occur (e.g., Metalsky, Joiner, Hardin, & Abramson, 1993). Although research has yielded data consistent with this hypothesis, in one study in which interparental conflict was conceptualized as the stressor (O'Donnell, Moreau, Cardemil, & Pollastri, 2010), a pessimistic attributional style reduced children's risk for depressive symptoms in response to interparental conflict. Although this result was not expected and is somewhat difficult to interpret, it helps illustrate that the integration and application of broad theories on the development of psychopathology might enrich our understanding of the association between interparental conflict and child adjustment problems. It should also be acknowledged that children's threat and self-blame appraisals of their parents' conflict typically are thought of as mediators of the link between interparental conflict and child adjustment problems, but several investigators have also conceptualized them as moderators. This conceptualization is similar to that described earlier for theories of depression, in which children are proposed to have cognitive predispositions toward

interpreting and explaining negative events in a particular manner. With threat and self-blame appraisals, children may have characteristic ways of viewing interparental conflict or conflictual interactions in general (e.g., as potentially escalating and spilling over into other interactions), and this cognitive style might moderate the link between interparental conflict and child adjustment problems. A study by Kerig (1998) found some evidence of this: Children's self-blame and threat appraisals potentiated some of the associations examined between interparental conflict and child adjustment problems. However, others have failed to find such moderation effects (e.g., Grych, Fincham, McDonald, & Jouriles, 2000).

Neurobiological Responses to Interparental Conflict (Stress) Considering child attributes in general, and biological factors such as children's physiological responses to stress specifically, will likely help explain the multifinality of child outcomes in the context of interparental conflict and offer more precision in delineating pathways to specific child outcomes. Researchers have noted that physiological reactivity, regulation, and recovery in response to stress may modify the type and severity of child adjustment problems that develop in the context of interparental conflict (El-Sheikh & Erath, 2011; Steinberg & Avenevoli, 2000). Biological responses to stress may serve as a vulnerability (i.e., risk) factor to exacerbate relations between interparental conflict and child adjustment, or may serve as a protective factor to buffer children from the negative consequences of exposure to interparental conflict. Exposure to stress elicits responses from the autonomic nervous system (ANS) and the HPA axis. Children's ANS functioning has more commonly been examined as a moderator of relations between interparental conflict and child adjustment problems, whereas HPA axis functioning (e.g., cortisol reactivity, recovery, or diurnal patterns) has more commonly been tested as a mediator. This is consistent with direct-effect models based on trauma theory (described earlier), which posit that exposure to interparental conflict predicts changes in HPA axis functioning, which in turn is responsible for later adjustment problems (e.g., Davies, Sturge-Apple, & Cicchetti, 2011; Davies, Sturge-Apple, Cicchetti, Manning, & Zale, 2009; Koss et al., 2013; Lucas-Thompson, 2012; see also Repetti, Robles & Reynolds, 2011, for theoretical discussion of risky families and HPA axis functioning). This section focuses on research that documents children's ANS activity as a moderator of relations between interparental conflict and child adjustment. The ANS is responsible for involuntary functions, such as cardiovascular functioning and respiration, and is involved in the human stress response. El-Sheikh and Erath (2011) outlined five ways that ANS activity may increase risk for, or protect against, children's adjustment problems in the context of family conflict: (1) Baseline levels of ANS activity may affect how ready children are to respond to interparental conflict. In addition, ANS activity may (2) indicate how sensitive children are to the effects of interparental conflict, (3) influence children's emotions, cognitions, and behavior in response to interparental conflict in the short term, (4) influence activity of other stress response systems, like the HPA axis, and (5) change the response of other systems over time so that they become dysregulated.

The ANS is divided into two branches, the parasympathetic nervous system (PNS) and the SNS. Both are considered to be stable, individual difference variables, supporting their role as moderators of relations as opposed to mediators. For example, indicators of PNS (El-Sheikh, 2005; Hinnant, Elmore-Staton, & El-Sheikh, 2011) and SNS (El-Sheikh, 2007; Hinnant et al., 2011) activity have been found to be stable among school-age and early adolescent children over a 2- to 3-year time period. However, ANS functioning in infancy and early childhood may be more malleable and, therefore, less likely to serve as a moderator during those developmental periods (El-Sheikh & Erath, 2011). The PNS supports resting functions and serves to decrease heart rate, dilate blood vessels, and increase digestive and reproductive efforts. Reduced PNS activity, or vagal withdrawal (also referred to as vagal suppression), is an automatic response to stress. It allows diversion of resources to deal with the stressor, including engaging attention, gathering and processing information, and using appropriate response strategies. Physiologically, reduced PNS activity results in rapid, moderate increases in heart rate. Once the stressor has diminished, PNS activity increases again to return the organism to a resting state. In general, vagal withdrawal is an adaptive response to stress and is related to positive outcomes (Beauchaine, 2001, 2012; Calkins & Keane, 2004; Porges, 2007). Extreme vagal withdrawal, however, or vagal augmentation (increased PNS activity during stress) is considered a maladaptive response to stress and is related to negative physical and mental health outcomes (Beauchaine, 2012; Bornstein & Suess, 2000; Calkins, Graziano, & Keane, 2007). Respiratory sinus arrhythmia (RSA), which is the rhythmic fluctuation in heart rate during phases of the breathing cycle, is used as a measure of PNS activity. Baseline levels of RSA are used to index vagal tone and RSA change from baseline during stress reflects PNS reactivity. In the context of interparental conflict, research has found that low baseline RSA is a vulnerability factor; children with low levels of baseline RSA have greater internalizing, externalizing, and physical health problems (e.g., Dietrich et al., 2007; El-Sheikh, Harger, & Whitson, 2001; Katz & Gottman, 1995). Decreased RSA activity, indicative of PNS withdrawal, is protective, whereas increased RSA activity (or less withdrawal) has been found to exacerbate the negative relation between interparental conflict and child adjustment problems (El-Sheikh & Whitson, 2006; Katz, 2007; Leary & Katz, 2004). For example, ElSheikh and Whitson (2006) found an interaction between vagal withdrawal and marital conflict, such that for girls, vagal withdrawal to a lab stressor was associated with fewer internalizing problems 2 years later in the context of high interparental conflict. Notably, although most research has focused on school-age and older children, G. A. Moore (2010) found that interparental conflict was related to lower baseline RSA and less RSA withdrawal among 6-month-olds during a still-face paradigm. This study, however, did not test RSA and RSA reactivity as a moderator of relations between interparental conflict and infant adjustment. Activation of the SNS during stress prepares the body for the flight-or-fight response, resulting in increased heart rate, blood pressure, and oxygen flow throughout the body and mobilizing the body to respond to stress. In the short term, SNS activation is adaptive; however, chronic activation may lead to “wear and tear” on the body, resulting in physical and mental health

problems. Noninvasive measures of the SNS include skin conductance level (SCL), and preejection period (PEP), which measures the time between heart beat onset and the ejection of blood into the aorta. Changes from baseline in SCL and PEP during stress or a challenge task assess SCL reactivity and PEP reactivity, respectively. Increases in SCL and decreases in PEP (i.e., shorter intervals) as compared to baseline levels reflect greater SNS activity, and are associated with increases in heart rate. In the context of interparental conflict, children who exhibit high SCL reactivity have higher levels of adjustment problems, particularly among girls (El-Sheikh, 2005; El-Sheikh, Keller, & Erath, 2007). Fewer studies have examined PEP as a moderator in the context of interparental conflict. Obradović, Bush, and Boyce (2011) found that children who showed decreased PEP from baseline during a task (indicating increased SNS activity, and referred to as high PEP reactivity in this study) had elevated internalizing problems in the context of high marital conflict as compared to children who showed increases in PEP during a challenge task. Most of the previous research has examined PNS and SNS functioning as separate moderators, but recent theoretical perspectives advocate examining interactions between PNS and SNS functioning (Berntson & Cacioppo, 2004; Del Guidice, Hinnant, Ellis, & El-Sheikh, 2012; ElSheikh & Erath, 2011). The PNS and SNS do not operate in isolation, and interaction models may better capture children's stress responses. Children's risk for adjustment problems in the context of interparental conflict, therefore, may not rest solely on how one of these systems operates but rather on whether these two systems work in tandem to promote the same directional response. For example, the adaptive calibration model (Del Guidice, Ellis, & Shirtcliff, 2011) posits four prototypical response patterns to stress, some more adaptive than others. The response patterns are sensitive, buffered, vigilant, and unemotional. Berntson and Cacioppo's (2004) doctrine of autonomic space also proposed four response patterns, which clearly map onto measurements of PNS and SNS cardiac-related activity. These include reciprocal parasympathetic activation (PNS activation, SNS inhibition), reciprocal sympathetic activation (SNS activation, PNS inhibition), coactivation (PNS activation, SNS activation), and coinhibition (PNS inhibition, SNS inhibition). Salomon, Matthews, and Allen (2000) found that reciprocal sympathetic activation is a normative response to stress in children. In three independent cross-sectional samples, El-Sheikh and colleagues (2009) examined interactions between PNS activity, SNS activity, and interparental conflict in the prediction of children's externalizing behavior problems. The results were consistent across samples and reporters (parents, teachers): Coactivation and coinhibition of the PNS and SNS placed children at increased risk for externalizing problem in the context of marital conflict. In contrast, reciprocal parasympathetic and reciprocal sympathetic activation operated as a protective factor. The results suggested that coordinated actions of the PNS and SNS buffer children from the adverse effects of interparental conflict. In a longitudinal study, El-Sheikh, Keiley, Erath, and Dyer (2013) also found evidence that coinhibition may be a risk factor for poor adjustment. Children who showed coinhibition to a laboratory challenge task had increasing trajectories of depressive symptoms over a 3-year period in the context of interparental conflict. However, they also found that reciprocal parasympathetic activation was

related to elevations in depressive symptoms, which is inconsistent with El-Sheikh et al. (2009). Interactions may occur not only between PNS and SNS functioning but also between baseline and reactivity levels within the same system. For example, Hinnant and El-Sheikh (2009) found that interactions among interparental conflict, baseline level of RSA, and RSA reactivity, such that low baseline levels of RSA in the context of higher RSA suppression strengthened the positive association between marital conflict and internalizing problems 2 years later, whereas low baseline levels of RSA in the context of RSA augmentation (increases in RSA activity) predicted greater externalizing problems 2 years later for children experiencing higher levels of interparental conflict. In other words, although low basal levels of RSA are related to greater child adjustment problems, RSA regulation (withdrawal/suppression versus augmentation) may buffer children from experiencing problems or further exacerbate child adjustment problems in the context of interparental conflict. Finally, although HPA axis activity can be conceptualized as a mediator of relations between interparental conflict and child adjustment, there is some evidence that it may interact with PNS reactivity to predict child adjustment problems. Specifically, across two samples, ElSheikh, Arsiwalla, Hinnant, and Erath (2011) found that children with the lowest levels of anxiety and depression symptoms showed a pattern of higher basal cortisol levels and higher basal RSA. Future research may consider examining interactions between HPA and ANS functioning in the context of interparental conflict in predicting children's adjustment. Alternatively, ANS functioning may alter the mediating role of the HPA axis in the interparental conflict–child adjustment link (i.e., moderated mediation). The role of children's biological responses to stress as moderators may be more complex than previous research has documented. Whereas the research discussed earlier shows that low basal PNS activity, less PNS withdrawal or PNS augmentation, and higher SNS reactivity are risk factors, exacerbating the risk of child adjustment problems in the context of interparental conflict, recent perspectives in the literature call for research to move beyond diathesis–stress or “dual-risk” models (Belsky & Pluess, 2009; Ellis, Boyce, Belsky, Bakermans-Kranenburg, & Van Ijzendoorn, 2011). Based on developmental-evolutionary models, biological sensitivity to context theory (Boyce & Ellis, 2005; Ellis, Essex, & Boyce, 2005) and differential susceptibility to environmental influences theory (Belsky, 1997; Belsky & Pluess, 2009) propose that the biological factors, including neurobiological factors (e.g., pattern of physiological responding to stress), that place children at risk for adjustment problems in the context of adverse environments are the same factors that allow children to thrive in positive environments. For example, biological sensitivity to context theory would predict that there is a curvilinear relation between heightened ANS functioning (as well as adrenocortical and other stress responses) and interparental conflict in predicting children's adjustment problems. That is, whereas children with increased RSA or SNS reactivity may be at increased risk for adjustment problems in the context of destructive interparental conflict, they may also be more sensitive to the positive effects of secure, stable, and positive interparental relationships. Differential susceptibility to environmental influences theory suggests further that there will be no association between destructive interparental conflict and child adjustment among children

who do not have heightened RSA or SNS reactivity.

Genetic Influences A goal of behavioral genetics research is to tease apart the relative contributions of genetic and environmental influences on behaviors. Recent behavioral genetics studies have focused on these sources of influence in the cascade of effects from interparental conflict through parenting to child adjustment. In an elegantly designed study, Harold and colleagues (2013) found compelling evidence of environmental, rather than genetic, influences for the relation between interparental conflict and hostile parenting and for the mediating role of hostile parenting in the relation between interparental conflict and children's externalizing problems. Moreover, across varying levels of parents' genetic relatedness to their children (i.e., neither parent, mother only, father only, or both parents genetically related to the child), they found the magnitude of the effect of interparental conflict on hostile parenting to be stronger for fathers than for mothers. Finding greater effects of interparental conflict on fathering than mothering converges with findings from other studies on interparental conflict and parenting (e.g., Davies, Sturge-Apple, Woitach, & Cummings, 2009). In a related study, Harold, Elam, Lewis, Rice, and Thapar (2012) examined links among interparental conflict, hostile parenting, and parent and child antisocial behavior. They found fathers' parenting and fathers' antisocial behavior predicted child antisocial behavior indirectly via fathers' parenting in genetically related and genetically unrelated father-child pairs. For mothers, however, only mothers' antisocial behavior mediated the link between interparental conflict and child antisocial behavior in genetically unrelated mother-child pairs. In genetically related pairs, both mothers' parenting and mothers' antisocial behavior mediated the link. Studies such as this, which consider multiple pathways of influence, will be important for articulating how, and how much, the effect of interparental conflict on child adjustment differs across mothers and fathers. The design of the Harold et al. (2013) study controlled for the effects of passive genotypeenvironment correlation (i.e., shared genetic factors underlie and thus influence parent and child behavior) on the links between interparental conflict, parenting, and child adjustment. It did not examine evocative genotype-environment correlations—that is, whether genetically influenced child behaviors may have evoked particular parenting responses (i.e., commonly referred to as child effects). A study examining whether interparental conflict moderates evocative child effects on parenting found evidence for genetic as well as environmental child evocative effects and also found that those effects are moderated by interparental conflict (Ulbricht et al., 2013). The study employed a large sample that included monozygotic twin pairs; dizygotic twin pairs; sibling pairs from nondivorced families; and sibling, half-sibling, and unrelated sibling pairs from stepfamilies, allowing careful parsing of genetic and environmental effects. Child-based genetic effects, along with shared and nonshared environmental (child) factors were associated with mothers' and fathers' negative parenting, and these effects varied across levels of interparental conflict. Specifically, as marital adjustment declined, evocative child effects on parenting increased, while the role of shared family experiences declined. However, there were some differences in the pattern of specific

relations of marital adjustment variables (e.g., global maladjustment, conflict over children) on child-based genetic and child-specific nonshared environmental effects on parenting across mothers and fathers. As interparental conflict increased, fathers' negative parenting was more influenced by nonshared environmental effects whereas mothers' negative parenting was more influenced by their children's genetically mediated characteristics. Assuming the commonly accepted idea of parenting serving to mediate the relation between interparental conflict on child adjustment, findings from this study are consistent with a dynamic process in which both interparental conflict and individual characteristics of the child affect parenting, which in turn affects child adjustment. What this study adds is that child effects on parenting are stronger in families with greater interparental conflict. In contrast to behavioral genetics research, molecular genetics focuses on understanding effects of particular genes and genetic configurations on behavior. Molecular genetics studies have shed light on several aspects of the link between interparental conflict and child adjustment. For example, Sturge-Apple, Cicchetti, Davies, and Suor (2012) examined whether differential susceptibility to environmental influences (see Belsky, 1997 for details of this theory) might explain variation in the link between interparental conflict and parenting sensitivity. In this study, investigators examined oxytocin receptor (OXTR) and serotonin transporter (5-HTT) genes—both of which have been previously linked (albeit not in consistent ways) to parenting (Bakermans-Kranenburg & van IJzendoorn, 2008; Mileva-Seitz et al., 2011)—as moderators of the relation between interparental conflict and mothers' parenting sensitivity. Results were consistent with differential susceptibility theory. For OXTR rs53576, mothers with the GG genotype showed greater differential maternal sensitivity at different levels of interparental conflict; for the 5-HTT promoter polymorphism, mothers with one or two copies of the short allele demonstrated differential susceptibility for both sensitive and harsh parenting. In a study examining etiological pathways to two child disorders that often covary (attentiondeficit/hyperactivity disorder [ADHD] and oppositional defiant disorder [ODD]), Martel and colleagues (2012) found moderating effects of the 5-HTT promoter polymorphism on the relation between children's self-blame for interparental conflict and symptoms of ODD. Children with the short allele of the 5-HTT promoter polymorphism experienced greater ODD symptomatology regardless of their level of self-blame for interparental conflict, whereas those without the short allele had greater ODD symptomatology when self-blame for interparental conflict was higher. The studies reviewed in this section serve as examples to help illustrate the complexity and interplay of the biological, psychological, and social processes that govern the ways that family conflict and parenting shape child adjustment, and vice-versa. Research using genetically sensitive designs, such as those described, and careful consideration of the issues raised by these studies, such as genetic and environmental sources of influence, shared versus nonshared environmental influence, the influence of mothers' versus fathers' parenting, the directionality of effects (e.g., child evocative effects), and the precise nature of the effects (e.g., direct, mediated, moderated, and moderated effects consistent with differential susceptibility), can greatly sharpen theory and knowledge on the variability in outcomes among children exposed to interparental conflict.

Interparental Conflict and the Broader Family and Community Context Children are embedded in their families, which are in turn embedded in a complex network that includes peer groups, neighborhoods, and broader cultural contexts. Cumulative risk models of child psychopathology highlight the potential importance of considering these different contexts when attempting to understand the association between interparental conflict and child adjustment. Specifically, destructive interparental conflict commonly occurs in family contexts in which there are also other types of family problems (e.g., parental psychopathology, parental criminal behavior, economic stressors, etc.). Cumulative risk models suggest that it is the accumulation of family problems, rather than the presence of any single type of family problem, that is responsible for child adjustment problems. In other words, this perspective suggests that destructive interparental conflict may not be sufficient to produce child adjustment problems, especially if it occurs alone or in isolation from other family problems. However, destructive interparental conflict, in combination with other family problems, is hypothesized to be much more likely to lead to child adjustment problems. Quite a bit of empirical evidence has amassed that is consistent with certain facets of these cumulative risk models of child adjustment problems, and many of these studies have included interparental conflict as one of the risk factors. For example, Rutter (1979) examined the influence of multiple risk factors for child adjustment problems (e.g., severe marital discord, paternal criminality, maternal mental disorder) and found that the accumulation of risk factors had an exponential association with child adjustment problems. Specifically, the presence of an isolated family stressor did not increase risk for child adjustment problems; in contrast, the presence of two family stressors resulted in a fourfold increase in child adjustment problems, and the presence of four or more family stressors resulted in a tenfold increase in child adjustment problems. Appleyard, Egeland, van Dulmen, and Sroufe (2005) similarly examined the influence of multiple risk factors (e.g., IPV, child maltreatment, low SES) for child adjustment problems, and these investigators found evidence for a linear relation between the number of risk factors present and child adjustment problems, but no evidence for the exponential association. In this section of the chapter, we focus on documented interactions between interparental conflict and other variables in the prediction of child adjustment problems. That is, we focus on family and community variables that have been found to moderate associations between interparental conflict and child adjustment problems. These variables include direct associations between interparental conflict and child adjustment problems as well as indirect associations involving mediating pathways (moderated mediation). Although we cover variables that have been found to amplify risk for child adjustment problems, we emphasize those that have been found to mitigate risk. We organize this section around three topics: attributes of the family, peer relationships, and community attributes. Attributes of the Family A commonly advanced hypothesis is that positive aspects of the parent-child relationship

protect children from the adverse effects of interparental conflict. Consistent with this, several investigators have found positive parenting behaviors or a positive parent-child relationship to buffer children from the adverse effects of destructive interparental conflict. Frosch and Mangelsdorf (2001), for example, found that observers reported more child behavior problems when the children's parents exhibited higher levels of interparental conflict and mothers were less warm and supportive in their parenting. However, if mothers were more warm and supportive in their parenting, fewer child behavior problems were reported, even in the presence of higher levels of interparental conflict. Similarly, others have found a positive parent-child relationship to weaken the association between destructive interparental conflict and child adjustment problems (e.g., Davies, Harold, Goeke-Morey, & Cummings, 2002; Kennedy, Bybee, Sullivan, & Greeson, 2010; Skopp, McDonald, Jouriles, & Rosenfield, 2007). In addition to finding that parental warmth and positive parental behaviors (e.g., affection, engagement, positive structuring, and responsiveness) mitigate the association between interparental conflict and a variety of negative child outcomes, Katz and Gottman (1997) found that parental “emotion coaching” (acting as a coach for the child during their child's stressful times and emotional moments—teaching the child to identify and regulate emotions) reduced the magnitude of the association between interparental conflict and child adjustment. Katz, Hunter, and Klowden (2008) replicated this, finding that mothers' emotion coaching reduced the magnitude of the association between IPV and children's odd behavior during a peer provocation. Several investigators have attempted to examine how parents' conversations with their children about interparental conflict might influence child adjustment problems. McDonald, Jouriles, Rosenfield, and Leahy (2011) examined mothers' responses to children's questions about interparental conflict in a sample of families in which there was a recent incident of IPV. They found that the extent to which mothers' responses addressed the explicit content of their children's questions about interparental conflict was negatively associated with child adjustment problems. This finding is consistent with those reviewed in the preceding paragraph, suggesting that positive aspects of the parent-child relationship may buffer children from the adverse effects of destructive interparental conflict. Gomulak-Cavicchio, Davies, and Cummings (2006) also examined mothers' communications with children about their disputes, but these investigators found that maternal communications highlighting family stability and warmth appeared to magnify the risk posed by interparental conflict. These authors were prudent in their interpretation of this counterintuitive finding, but the finding is not inconsistent with the finding of McDonald et al. That is, it may be that answering the child's specific questions, rather than offering general comfort or a sense of safety, may help a child better cope with interparental conflict. Collectively, findings on the effects of positive parent-child communications and/or positive parent-child relationships suggest that they are likely to be more complex than a simple buffering or protective hypothesis might suggest. To offer an example of this complexity, we highlight a finding in a study conducted by Skopp, McDonald, Jouriles, and Rosenfield (2007). In this study, parental warmth toward the child (operationalized by items such as “I comfort my child when he/she seems scared, upset or unsure” and “My child and I hug and/or kiss each other”) was examined as a moderator of the

association between men's IPV and child externalizing problems. The findings were complex and differed for mothers' and fathers' warmth expressed toward male and female children. However, one of the more interesting findings was that men's IPV related positively to child externalizing problems in families where the men were warmer toward the child. This finding is generally consistent with the finding reported by Gomulak-Cavicchio, Davies, and Cummings (2006), in that a presumably positive parent-child interaction appeared to amplify the risk associated with interparental conflict. It is also consistent with findings in other literatures suggesting that the risk for child conduct problems is amplified when antisocial fathers are highly involved in their children's care (Jaffee, Moffitt, Caspi, & Taylor, 2003) as well as with the social cognitive theory explanation that children are more likely to emulate an adult's behavior when they share a warm bond with that adult (Bandura, 1986). Again, the effects of positive parent-child communications and/or positive parent-child relationships appear to be complex. Parents' relationships with others in the family may also have an impact on how interparental conflict affects child adjustment, and in equally complex ways. Fosco and Grych (2007) assessed how often parents expressed positive affect toward family members (the specific target of the positive affect was not specified) and found that the association between interparental conflict and child adjustment problems was slightly stronger for families in which the parents expressed greater positive affect. In other words, more frequent expression of parental positive affect in the family appeared to exacerbate rather than buffer against the effect of interparental conflict on child adjustment. Fosco and Grych were appropriately cautious in interpreting this result. Nonetheless, this counterintuitive finding points to the inherent complexity in how family interactions transact with interparental conflict in the prediction of child adjustment problems. Maternal symptoms of psychopathology have also been found to moderate associations between children's exposure to interparental conflict and child adjustment problems. Specifically, maternal depressive symptoms and psychological distress amplify associations between children's exposure to IPV and child adjustment problems of preschoolers (MartinezTorteya, Bogat, von Eye, & Levendosky, 2009), school-age children (Levendosky & GrahamBermann, 2000), and adolescents (Levendosky, Huth-Bocks, & Semel, 2002). Levendosky and colleagues (2002) have offered several hypotheses for children's increased vulnerability to the effects of IPV in the context of maternal depression. These explanations involve other family processes, such as mothers' diminished emotional availability to their children when experiencing both IPV and depression, and children having to take on more of a caregiver role within the family. High rates of co-occurrence for IPV and physical child abuse (Appel & Holden, 1998; Jouriles et al., 2008) have sparked investigations of the joint effects of these two forms of family violence on child functioning. Although the results of these studies are not entirely consistent, most investigators report evidence for additive effects. That is, both forms of family violence appear to contribute independently to subsequent child adjustment problems (Jouriles, Mueller, Rosenfield, McDonald, & Dodson, 2012; Litrownik, Newton, Hunter, English, & Everson, 2003). Attempts to document multiplicative effects (e.g., show that the presence of child abuse

amplifies associations between IPV and child adjustment problems) have generally not been successful (Litrownik et al., 2003). Peer Relationships An important task during childhood is the development of beneficial and satisfying peer relationships. Acquiring the ability to develop and maintain meaningful peer relationships is among the most important developmental tasks during middle childhood and adolescence, and peer problems, although not technically a psychological disorder, might rightfully be conceptualized as a type of child adjustment problem. In fact, peer problems are characteristic of a wide range of child psychological disorders. As one might expect from the research reviewed thus far, interparental conflict has consistently been found to be associated with children's problematic peer relationships (e.g., Erel et al., 1998; Gottman & Katz, 1989; McHale et al., 1999; Stocker, Ahmed, & Stall, 1997; Stocker & Youngblade, 1999). When considering children's peer relationships in the link between interparental conflict and child adjustment problems, the question becomes whether peer relationships can influence what is arguably a within-family process. Wasserstein and LaGreca (1996) found that support from a close friend moderated the link between interparental conflict and grade-school children's adjustment problems; interparental conflict was associated with child problems only among children who had received low levels of support from a close friend. Rogers and Holmbeck (1997) found that a measure of peer avoidance (interpreted as perceived peer availability) moderated the association between IPV and child adjustment problems, so that children in high-conflict families who reported greater peer availability had fewer adjustment problems than their counterparts in high-conflict families who reported less peer availability. Criss, Pettit, Bates, Dodge, and Lapp (2002) reported a similar moderator effect in a longitudinal study with a sample of younger children (kindergarten to second grade). Specifically, marital conflict was associated with externalizing problems only among children with low levels of peer acceptance. Community Families, of course, are nested within the broader context of communities. There has been speculation, but little research, on whether children's exposure to interparental conflict and violence in the family interacts with aspects of the family's community in influencing child development (Cicchetti & Lynch, 1993). On the other hand, the concept of polyvictimization— experiencing multiple types of violence across multiple settings— has begun to attract considerable attention among child victimization researchers (Finkelhor, Ormrod, & Turner, 2007). This is because it is becoming increasingly clear that exposure to multiple types of violence (e.g., violence between the parents, violence within the community, violence at school) is much more likely than exposure to a single, isolated type (Margolin et al., 2009). In general, research on polyvictimization suggests that children who are exposed to multiple types of violence fare worse than those exposed to a single type of violence. For example, Mrug, Loosier, and Windle (2008) found that cumulative violence exposure across settings (home, school, and community) was a stronger correlate of adolescents' externalizing and

internalizing problems than exposure in any single setting. Similarly, Margolin, Vickerman, Oliver, and Gordis (2010) found that an index of children's cumulative violence exposure (including IPV, community violence, and parent-to-youth aggression) was a stronger predictor of youth problems than violence exposure in any single domain. What is less clear is how these different types of violence exposure might interact with one another in predicting child adjustment problems. Investigations of how children's exposure to different types of violence interact with one another often provide no evidence of interaction effects (e.g., Kennedy, Bybee, Sullivan, & Greeson, 2010), or interactive effects that are contrary to expectations. For example, in a prospective study, high levels of community violence exposure mitigated relations between witnessing IPV and adolescents' anxiety, depression, and delinquency (Mrug & Windle, 2010). Possible explanations for this result include the idea that IPV might simply be less salient to (and perceived as a more normal phenomenon for) adolescents whose environments also include community violence. There are a number of studies on how poverty or low family SES might moderate the association between interparental conflict and child problems. As noted, cumulative stress models of child psychopathology (e.g., Rutter, 1979) suggest that living in poverty might potentiate associations between interparental conflict and child adjustment problems, and findings from early studies were consistent with this hypothesis. For example, Jouriles, Bourg, and Farris (1991) found that marital arguments were more strongly associated with child conduct problems in families of lower, rather than higher, SES. However, more recent research suggests that the joint effects of these two family stressors may not be that clear cut. El-Sheikh and colleagues (2008) examined SES as a moderator of the emotional security theory explanation for the link between interparental conflict and child adjustment problems, and they found no differences based on high or low SES. Yoo and Huang (2012) found that poverty moderated associations between IPV and child adjustment problems, but the magnitude of the association was smaller among impoverished families. Culture Many of the associations between specific forms of interparental conflict and child adjustment problems have been documented in different countries and cultural contexts. Similarly, associations between hypothesized mediators of the relation between interparental conflict and child problems have been linked to child problems in different cultural contexts. To offer some concrete examples, the longitudinal association between witnessing parental IPV during childhood or adolescence and being involved in a violent relationship as an adolescent or young adult has been documented in many different samples within the United States (e.g., Amato & McNeal, 1998; Cui, Durtschi, Donnellan, Lorenz, & Conger, 2010; Ehrensaft, Cohen, & Brown, 2003; Ireland & Smith, 2009; Linder & Collins, 2005) as well as in samples in Canada (Wolfe, Wekerle, Scott, Straatman, & Grasley, 2004), New Zealand (Fergusson, Boden, & Horwood, 2006), and the Philippines (Fehringer & Hindin, 2009). Associations between children's appraisals of interparental conflict as threatening and child internalizing problems have been observed in the United States (e.g., Grych et al., 2000), Canada (Kerig, 1998), Wales (Davies, Harold, Goeke-Morey, & Cummings, 2002), and Australia (Dadds,

Atkinson, Turner, Blums, & Lendich, 1999). Bradford and colleagues (2003) found evidence for the hypothesis that interparental conflict affects child adjustment indirectly through disruptions to parenting in samples of adolescents from Bangladesh, Bosnia, China, Colombia, India, Germany, Palestine, South Africa, and the United States. Experimental studies of children's immediate responses to conflict between adults have been observed to be similar among United States and Israeli children (Shamir, Cummings, Davies, & Goeke-Morey, 2005) and among children from Welsh families (Davies et al., 2002; Goeke-Morey, Cummings, Harold, & Shelton, 2003). Children's emotional security as a mediator of relations between interparental conflict and child adjustment problems has been documented with families from the United States (e.g., Cummings, Schermerhorn, Davies, Goeke-Morey, & Cummings, 2006) and Wales (Davies et al., 2002). In short, it appears that many of the documented relations between interparental conflict and child adjustment problems are not specific to any particular culture. What is much less clear is how specific aspects or dimensions of one's culture might influence associations between interparental conflict and child adjustment problems. In general, there is a dearth of theory specific to this topic, with several notable exceptions. For example, McLloyd and colleagues (2002) have discussed the potential importance of extended family networks and have suggested that relations between interparental conflict and child adjustment problems may be weaker among ethnic minorities in the United States because of their extended family networks. Such extended family networks are hypothesized to provide support to family members and presumably buffer children from the negative impact of conflict exposure. Empirical support for this particular hypothesis is mixed, with some researchers finding results that could be interpreted as consistent with this argument (e.g., Tschann, Flores, Pasch, & VanOss Marin, 1999) and others not (e.g., El-Sheikh et al., 2008). Developing sound theoretical explanations for possible differences across cultures and then examining cultural differences in the association between interparental conflict and child adjustment remains an important direction for future research.

Translational Implications For many, translational research refers to the formation of clinical interventions through the use of scientific knowledge. Applied to the topic of this chapter, translational implications might be conceptualized to refer to the ways in which knowledge about the association between interparental conflict and child adjustment might be used to help promote healthy child development. Of course, one simple answer to this question is preventing children's exposure to destructive interparental conflict. Doing this might involve the deployment of intervention programs targeted at couples with children, such as programs that attempt to develop healthy and effective conflict resolution skills. However, preventing or halting destructive interparental conflict is not always simple or easy. Even the severing of a couple's intimate relationship (divorce, separation) does not always result in children no longer being exposed to destructive interparental conflict. Furthermore, what should be done once children have already been exposed to destructive interparental conflict?

This section starts with a review of empirically supported interventions for children who have been exposed to interparental conflict. We begin with those interventions designed specifically for children exposed to interparental conflict; most of these interventions focus on children who have been exposed to IPV. Next, we review interventions that focus on reducing destructive interparental conflict and interventions that are salient to children's exposure to interparental conflict but that do not focus specifically on children exposed to violence. Our review of interventions is comprehensive, but it is not intended to be exhaustive. We hope that it is illustrative, however, providing readers with an idea of the breadth of interventions pertinent to this topic as well as ideas for further intervention work in this area. Figure 12.1 summarizes the various mechanisms and pathways by which interparental conflict is believed to influence child adjustment. This diagram summarizes the theory and data reviewed in earlier sections and provides an organizing framework for discussing the translational implications of research on children's exposure to interparental conflict.

Figure 12.1 Organizational Framework for Mechanisms and Pathways by which Interparental Conflict and Violence are Theorized to Influence Child Adjustment.

Interventions Designed for Children Exposed to IPV Project Support Project Support (Jouriles et al., 2001, 2009; McDonald, Jouriles, & Minze, 2011) was originally designed to reduce conduct problems among children, ages 3 to 8 years, who had been exposed to frequent and severe IPV. Conduct problems are frequently observed among children who have been exposed to such violence, and they tend to be fairly stable without intervention. The Project Support intervention is based on social cognitive theory on the development of antisocial behavior, basic research linking interparental conflict to parenting, research on dimensions of parenting that moderate associations between interparental conflict and child adjustment problems, and clinical research evaluating interventions for child conduct problems. It is consistent with principles of developmental psychopathology, which point to the parent-child relationship as a primary source of influence on child development. Project Support includes two components, the provision of social and instrumental support to the mothers and parent training, which is conceptualized as the principal avenue through which children's conduct problems are ameliorated. The parent training component attempts to improve the quality of the parent-child relationship by teaching and encouraging warm, attentive, and supportive parenting behaviors. It also attempts to reduce children's exposure to hostile and aggressive behavior and to create a more predictable environment for children. Intervention services are delivered in weekly home visits for about 6 months. Two randomized controlled trials of Project Support (Jouriles et al., 2001, 2009; McDonald, Jouriles, & Skopp, 2006) have been conducted with families in which the mothers had recently departed from a domestic violence shelter and had a child who met diagnostic criteria for oppositional defiant disorder or conduct disorder. Families in these studies were randomly assigned to receive Project Support or to a services-as-usual comparison condition. Results of both trials indicated that Project Support was effective in improving mothers' parenting and reducing children's conduct problems, and results observed at posttreatment were maintained over both studies' follow-up periods (20 months posttreatment in study 1, 12 months in study 2). In both studies, rates of clinical levels of child conduct problems at the last follow-up assessment were markedly lower in the Project Support condition than those in the services-as-usual condition (Jouriles et al., 2009; McDonald et al., 2006). In addition, mothers in the Project Support condition reported their children to be happier, to have better social relationships, and to have lower levels of internalizing problems relative to children in the comparison condition (McDonald et al., 2006). Mothers were less harsh and more consistent in their parenting, and mothers' parenting was found to mediate the effect of the intervention on children's conduct problems (Jouriles et al., 2009). Kids' Club The Kids' Club is a 10-week group program designed to address adjustment problems among 5- to 13-year-old children who have been exposed to IPV (Graham-Bermann, 1992). Grounded in social cognitive theory, the intervention focuses on teaching children about IPV, promoting

healthy beliefs and attitudes, reducing anxiety and worry, and improving social skills (GrahamBermann & Hughes, 2003). Two forms of the intervention were evaluated in a randomized controlled trial (n = 181) with a sample of mothers and their children ages 6 to 12: a group intervention offered only to the children, and one with the child group intervention plus a separate parenting group intervention for mothers. The comparison group received no services. For externalizing problems and attitudes about violence, only the child-plus-mother intervention resulted in positive change. Internalizing problems improved for children in all three groups. However, when clinical levels of internalizing problems were considered, only the intervention conditions reduced internalizing problems to nonclinical levels, with the childplus-mother condition showing a greater reduction (Graham-Bermann, 2000; GrahamBermann, Lynch, Banyard, Devoe, & Halabu, 2007). A Preschool Kids' Club, which targets anxiety reduction and conduct problems and which also includes a parenting group for mothers, has also been evaluated, although with somewhat less efficacious outcomes (Basu, Malone, Levendosky, & Dubay, 2009; Graham-Bermann & Halabu, 2004). Child-Parent Psychotherapy Child-Parent Psychotherapy (CPP) is an intensive intervention that focuses on improving the mother-child relationship to help children cope with trauma symptoms (Lieberman, Van Horn, & Ippen, 2005). CPP is based in attachment theory, social cognitive theory, and ecological models of therapy. It was designed for preschool-age children and focuses primarily on the mother-child dyad as the vehicle through which to help children exposed to IPV. Goals include promoting affect regulation in both the mother and child, changing maladaptive mother and child behavior as well as mother-child interaction, facilitating developmentally appropriate mother-child interaction, developing a trauma narrative, and developing hope and trust in the parent-child relationship (Lieberman, Ghosh Ippen, & Van Horn, 2006). Lieberman and colleagues (2006) conducted a randomized controlled trial evaluating CPP in comparison to monthly case management plus individual psychotherapy. Participants were 75 children ages 3 to 5 years who had witnessed IPV and their mothers. Families participated in an average of 32 sessions over the course of a 1-year treatment period. Results at posttreatment indicated that trauma symptoms and behavior problems decreased among children in the CPP condition but not among those in the comparison condition (Lieberman, Van Horn, & Ghosh Ippen, 2005). Improvements in behavior problems were maintained at the 6-month follow-up (results for trauma symptoms were not reported in the follow-up evaluation) (Lieberman et al., 2006). Community Advocacy Program Sullivan and colleagues developed and evaluated a community advocacy intervention for women who were victims of IPV. The program involved a trained advocate working individually with the women several times a week for 16 weeks to help the family access community resources and support. The intervention had been shown to reduce women's depression and the recurrence of violent victimization and to increase social support (Sullivan & Bybee, 1999). To help the children in violent families, the investigators expanded the intervention to include a support and educational group for children that focused on safety,

feelings, and self-respect. They evaluated this expanded intervention in a sample of 80 mothers and their children ages 7 to 11. Of the 9 hours per week that advocates spent with each family, approximately 5 were spent interacting with the children (Sullivan, Bybee, & Allen, 2002). The enhanced program yielded only modest effects on child well-being (Sullivan et al., 2002). Children in both groups experienced a decline in the amount of IPV witnessed, with this decline occurring earlier for children in the intervention condition. Children in both groups also had increased self-competence (self-concept and feelings of self-adequacy) at posttreatment, but by the 4-month follow-up, the self-competence of control-group children had returned to pre-intervention levels. There were no group differences in behavior problems. .

Trauma-Focused Cognitive Behavioral Therapy Trauma-Focused Cognitive Behavioral Therapy (TF-CBT; Cohen, Mannarino, & Deblinger, 2006) has been shown to reduce trauma symptoms in children who have developed posttraumatic stress disorder (PTSD) or symptoms of PTSD in response to potentially traumatizing events (e.g., child abuse). A randomized controlled trial evaluated its efficacy for 140 children ages 7 to 14 who had been exposed to IPV and who had IPV-related symptoms of PTSD (Cohen, Mannarino, & Iyengar, 2011). Children were randomly assigned to either a TFCBT condition or a child-centered therapy (CCT) condition. Children and their mothers were seen by the same therapist, but in separate sessions, once a week for 8 weeks, with the exception that in two of the TF-CBT sessions, the mother and child were seen together. TF-CBT provides education about trauma and teaches relaxation skills, affect expression and regulation, and coping skills. It also includes helping children develop a narrative about their IPV experience. In the course of developing the narrative, the therapist can assist children in correcting maladaptive cognitions and in coping with trauma reminders and events as they emerge. Children also share their experiences of IPV with their mother in the joint sessions. In CCT, children and parent direct the content of their sessions; the therapist provides reflection, empathy, emotional support, and encouragement. Results indicated that TF-CBT was superior to CCT in reducing trauma symptoms, particularly hyperarousal and avoidance symptoms, and anxiety symptoms. TF-CBT did not differ from CCT in reducing the reexperiencing of symptoms, which was not unexpected, given that the children lived in homes in which IPV recurred. Treatment effects were not observed for depression or behavior problems. A review of the collective evidence amassed for the effects of TF-CBT (i.e., not just for IPV-related trauma, but for other types as well, such as trauma related to child abuse) suggests that it appears to be efficacious in reducing trauma symptoms, but its efficacy for behavior problems and depression is less clear (de Arellano et al., 2014). Mapping to the Organizational Framework and Future Directions With the exception of Sullivan and colleagues' community advocacy intervention, each of the interventions just reviewed includes a component that targets changes in the parent-child relationship. Although there is some overlap, the theoretical foundation for the intervention varies across the interventions, as do the processes believed to mediate intervention effects. However, the parenting component to each appears more similar than different. Each focuses

on improving parenting skills and the affective quality of the parent-child relationship (in particular, improving maternal warmth and parent-child communication). Given the centrality of parents to the development of preschool and young school-age children, a focus on parenting is appropriate, and it may well be that the effects of these interventions are primarily attributable to their influence on parenting. However, the contribution of the other components of the interventions to improving child adjustment remains unclear. To illustrate, each of the interventions also targets reductions in maternal distress by providing emotional and social support and providing mothers with a safe arena in which to talk about their experience of IPV and how it has affected their lives and their children. However, changes in parenting were found to mediate the effects of Project Support on child behavior problems, whereas reductions in mothers' psychological distress did not. Furthermore, community advocacy, which successfully addressed mothers' emotional distress but which did not include a parenting component, did not have demonstrable effects on child behavior problems. This finding does not imply that additional intervention targets, such as responding to the suffering of mothers who have been victims of IPV, are not necessary or helpful (or simply morally right), but just that doing only that is not sufficient to help the children. Kids' Club, TF-CBT, and the community advocacy program include components targeting the child's immediate and delayed responses to IPV, including social cognitions and appraisals, negative affect, and affect regulation. Kids' Club and TF-CBT also work to help children develop coping skills to reduce the extent to which they are negatively affected by the conflict and IPV. Little is known about the relative efficacy of the parent-focused and child-focused components in those interventions that offer both. Graham-Bermann and colleagues' (2007) comparison of the child-only versus the child-plus-mother Kids' Club and Jouriles, McDonald, and colleagues' (2009) explicit analysis of mediating processes for Project Support reflect the kind of careful approach that is needed to optimize services for families of children exposed to IPV. A reasonable step for researchers would be to evaluate multiple plausible mediating pathways (such as those identified in the figure) simultaneously, from multiple plausible theoretical perspectives, to begin to drill down on the mechanisms that account most strongly for effects of IPV on child adjustment. Similarly, in conducting intervention research, it would be useful also to evaluate multiple mediating pathways, to help isolate those that account most strongly for intervention effects. It follows from several theories (e.g., attachment, cognitive-contextual, emotional security, trauma) that enhancing the child's sense of safety and security in the family should positively influence child mental health; however, the mechanisms responsible for this expected outcome differ across the different theories. Similarly, social cognitive theory and trauma theory posit different mechanisms to explain how IPV can lead to externalizing problems in children. Testing these mechanisms together in evaluation research will help clarify both theories and can shed light on how interventions based on these theories achieve their effects. Evaluating multiple mediating processes simultaneously would significantly advance knowledge in this area, and would aid intervention developers. Similarly, evaluating competing hypotheses about mediating processes and outcomes would help illuminate the key pathways by which IPV influences child adjustment. Finally, intervention researchers should attend to the multifinality

in outcomes of children exposed to IPV. As much as we need to know what outcomes can be prevented or improved, knowing that an intervention is not effective for a problem that occurs commonly among children in violent families can be just as important. As an example, the version of TF-CBT used for the IPV-exposed sample was not found to improve child behavior problems. However, in a different population (sexually abused children), TF-CBT developers found that a 16-week TF-CBT program that did not included a trauma narrative component helped improve parenting as well as child behavior problems (Deblinger, Mannarino, Cohen, Runyon, & Steer, 2011). The authors speculated that omitting the trauma narrative left more time to devote to teaching and practicing the parenting skills, thereby reducing behavior problems. This finding is consistent with what theory and data would suggest. Such efforts bring the field closer to being able to offer flexible and demonstrably effective interventions that can be tailored to meet a child's specific needs and circumstances. The interventions just reviewed were provided to children recruited from community agencies. The samples varied in the nature and levels of child adjustment problems, and the interventions differed in treatment modality (e.g., group versus dyadic) and duration (e.g., 10 weeks, 6 months, 1 year), yet the interventions were effective in helping the children. Indeed, after intervention, the problems of a sizable number of children were found to have diminished from clinical to normative levels on standard measures of child psychopathology. Thus, although the interventions were seemingly tightly focused in concept (i.e., targeting children “exposed to IPV”), they differ on a number of dimensions. The results from these interventions studies hold important implications for community agencies hoping to offer empirically supported interventions for children and their families. It suggests the need to assess child adjustment broadly and offer interventions that target those problems identified by the assessment. For now, this might mean that if trauma were the predominant problem, CPP or TF-CBT might be appropriate, depending on the child's age. If behavior problems predominate, Project Support or Kids' Club would be appropriate. Because comorbidity is the rule rather than the exception, however, many children in violent families may present with multiple, long-standing adjustment problems, including disrupted biological stress-response systems. Research thus far offers little guidance in this circumstance for prioritizing or organizing intervention approaches. In addition, many social services agencies are not equipped to offer a wide array of interventions, or lengthy interventions, which may be needed for families with severe and complex child adjustment problems. Research that helps to sort out how best to sequence or combine intervention components so that they can efficiently and effectively address multiple targets (such as the research on adaptations of TF-CBT) is needed. The field of implementation research has helped to underscore the importance of training practitioners to deliver interventions reliably and efficaciously. For psychosocial interventions, such training often involves education, behavioral practice, coaching/supervision until mastery, and follow-up performance monitoring (Fixsen, Naoom, Blase, Friedman, & Wallace, 2005). Such intensive training activities, however, are often beyond the means of most social services agencies (e.g., domestic violence agencies), especially those with high staff turnover rates, such as some child welfare agencies. Although there is pressure from the public policy arena for human services agencies to provide

evidence-based programs, there is also pressure, especially for publicly funded agencies, to provide services to as many people as possible. Such pressure may result in dollars being directed away from training and monitoring staff performance toward increasing the numbers of service providers and their caseloads. Such issues slow the uptake of effective intervention programming into the community. Another translational area that is underdeveloped is in the education of service providers. Intervention researchers are commonly embedded in doctoral programs in psychology or related fields, medical schools, or public health programs. However, most professionals who provide services to children in violent families have a master's degree, often in social work, counseling, or family therapy. Developing training partnerships with those educational pipelines might offer opportunities for building a stronger workforce and expanding its capacity.

Interventions to Reduce Interparental Conflict and Protect Child Well-being Happy Couples and Happy Kids Cummings, Faircloth, Mitchell, Cummings, and Schermerhorn (2008) conducted a randomized controlled trial evaluating the effects of Happy Couples and Happy Kids (HC/HK), a 4-week intervention that includes psychoeducation about marital conflict and its effects on children's emotional security in the interparental relationship and coaching in couple communication skills. A booster session was provided 6 and 12 months after the intervention. A community sample of families with a child between 4 and 8 years of age were randomly assigned to one of two forms of HC/HK or to a self-study comparison group. In one of the HC/HK groups, only the parents (the couple) participated; in the other HC/HK group, the child participated in a separate group while the parents were in session. Approximately half of the couples were maritally distressed at the outset of the study. Participants receiving the HC/HK intervention had superior outcomes in comparison to the self-study group, and outcomes did not differ across the two HC/HK conditions. For the HC/HK groups, destructive interparental conflict was reduced, constructive conflict increased (e.g., resolution, support behaviors), and there were improvements in parenting and child adjustment (Child Behavior Checklist total problems scale) at 1-year (Cummings et al., 2008) and 2-year (Faircloth, Schermerhorn, Mitchell, Cummings, & Cummings, 2011) follow-up assessments. Moreover, analyses of the patterns of change over time indicated that improvements in the couples' knowledge and handling of conflicts preceded improvements in marital satisfaction, parenting, and child adjustment outcomes. No changes were observed in the comparison group.

Interventions Salient to Children's Exposure to Interparental Conflict A number of interventions that were not originally designed to address the issue of children's exposure to interparental conflict have since been found to do so. These include marriage education and parenting interventions as well as various couples' interventions. We turn to these now.

Marriage Education and Parenting Interventions Over 30 years of research on marriage and relationship education programs attests to their utility in increasing couples' relationship satisfaction and improving the ways that couples manage conflict (Hawkins, Blanchard, Baldwin, & Fawcett, 2008). These effects can be very long-lived (e.g., 10 years; C. P. Cowan, Cowan, & Barry, 2011). There are many marriage and relationship education programs to choose from, several of which have been subjected to sustained and rigorous evaluation. Among these, however, thus far only a few have considered whether children benefit. Family Foundations is a “transition to parenthood” prevention program for expectant couples. It is designed to enhance coparenting, or the ways that parents work together, coordinate, support one another, and manage conflict about childrearing (Feinberg, Jones, Kan, & Goslin, 2010). The intervention includes education about the kinds of stresses and coparenting conflicts parents might expect after their baby is born and teaches couples communication, problem solving, and conflict- management skills. A sample of 169 couples was randomly assigned to participate in Family Foundations or to a no-treatment control condition. At the 3.5-year follow-up (when children were 3 years old), couples in the Family Foundations condition reported themselves to be less overreactive and less lax with their child, less likely to physically punish their child, and to have lower levels of parenting stress than couples in the comparison group. Family Foundations had positive effects on children's social competence; for families of boys, there were positive effects on couples' relationship quality and children's externalizing problems. There were no intervention effects on child internalizing problems. Based on knowledge that “the single most powerful predictor of fathers' engagement with their children is the quality of the men's relationship with the child's mother, regardless of whether the couple is married, divorced, separated, or never married,” P. A. Cowan, C. P. Cowan, and colleagues (2008, p. 54) developed a novel intervention designed to increase low-income fathers' engagement with their children (P. A. Cowan, C. P. Cowan, Pruett, Pruett, & Wong, 2009). In a randomized controlled trial of 289 families, expectant couples or those whose youngest child was under 7 years old were randomly assigned to one of three conditions: a 16week, father-only intervention; a 16-week couple intervention; or a 3-hour, single-visit comparison condition. The content of the father-only and couples group sessions was identical, and included sessions that focused on parenting and on couple issues. However, in-session exercises differed somewhat across the two formats (e.g., in the father-only group, some exercises that required the partner's participation were assigned as homework rather than completed during the session). At the 11-month follow-up, among families in the father-only and couple intervention conditions, fathers' engagement with their children increased, their parenting stress decreased, and fathers reported fewer parental conflicts about discipline. In addition, those in the couples' groups had stable couple relationship satisfaction over the follow-up period. This is in contrast to the comparison condition, where couple adjustment declined over the 11-month follow-up period. Post hoc tests indicated that children's adjustment remained stable in the two intervention conditions but declined in the comparison condition. This study is noteworthy for its focus on low-income, minority families (67% of the sample was Hispanic) and on fathers, in particular.

The Cowans and their colleagues also evaluated an intervention for parents with a child entering preschool (C. P. Cowan et al., 2011). The evaluation was a randomized controlled trial in which 100 families were assigned to a low-dose control condition (i.e., once-per-year consultation with the interviewer team) or to one of two 16-week intervention groups, one which focused on the couple relationship and one on parenting. Each group session included time allotted for presentation and discussion of the curricular material, and a 20- to 30-minute “check-in” period for couples to discuss issues that had arisen since the last session. The difference between the two group formats was how these check-in discussions were managed. In the couples group, the discussion focused on addressing the issues as a couple whereas in the parenting group, the focus was on addressing the issue from a parenting perspective. For example, if there were concerns about child discipline, the discussion in the couples group would center on how the couple could work together to resolve the issue. In the parentingfocused group, the discussion would focus instead on healthy discipline practices. In addition to contrasting these two intervention formats, the investigators examined their effects over a long period of time, conducting follow-up assessments when the children entered fourth grade and when they entered high school, using multimethod assessments of outcomes. Couple relationship satisfaction was stable over the 10-year follow-up period for those in the couples group, whereas it declined for those in the parenting and control groups (although there was a slight advantage for those in the parenting group compared to the control group). Couple communication quality improved through fourth grade for those in the couples group, whereas it was stable for those in the parenting group and declined for those in the control group. However, when the children entered high school, those effects had dissipated. There were also modest positive intervention effects on child adjustment, with child hyperactivity declining in the couples group condition but not in the control condition, and marginally significant findings of a declining trajectory for child aggression in the couples group but an increasing one in the control group. No group differences emerged for anxiety/depression. The results suggest that for primary prevention, a couple-focused program confers long-term protection for couples' and children's adjustment and that the protective effect is superior to those of an intervention focused solely on parenting. This line of research is exemplary in considering both couples and parenting interventions with sensitivity to the clinical setting and in evaluating treatment effects over a lengthy time period. Couples' Therapy In addition to increasing relationship satisfaction and reducing conflict generally, behavioral couples therapy has also been found to reduce interparental conflict over children (Gattis, Simpson, & Christensen, 2008). Moreover, in one evaluation of couples therapy that included assessment of child outcomes, improvements in child adjustment were observed during the treatment period; however, they were not maintained over the follow-up period (Gattis et al., 2008). Couples therapy effects on child adjustment were also examined in a randomized controlled trial of behavioral couples therapy (BCT) plus individual treatment for substance use in a sample of alcohol- and substance-using men involved in an intimate relationship (Kelley &

Fals-Stewart, 2002). The study included three groups. All participants received 20 weeks of individual treatment for substance use plus 12 additional weeks of intervention. For the additional 12 weeks, one group received couples therapy; one group received continued individual substance use treatment; and in one group, the couple attended substance abuse education sessions. Although the sample for the study was selected based on alcohol and substance use rather than on couple functioning, the mean level of relationship satisfaction for the couples that participated fell well into the distressed range. All three of the interventions deployed in the study were effective in reducing men's alcohol and substance use (although the couple intervention achieved somewhat greater reductions). However, in addition to its positive effects on alcohol and substance use, BCT also increased relationship satisfaction, reduced couple conflict and violence, and improved the psychosocial functioning of the children (children ranged from 6 to 16 years of age). Increases in relationship satisfaction and reductions in alcohol and drug use both contributed to the improvements in child adjustment. Specifically, at pretreatment, 50% of children of drug-using men and 32% of children of alcohol-using men exhibited clinical levels of adjustment problems. At the 12-month posttreatment follow-up, these figures were 15% and 5%, respectively. No statistically significant changes in couple relationship satisfaction or child adjustment were observed in either of the two comparison conditions at the 12-month assessment. A later pilot study evaluating treatment for alcohol-abusing men (Lam, Fals-Stewart, & Kelley, 2008) compared BCT to an intervention that included parent skills training plus BCT and to individual treatment for alcohol abuse. Although couples' adjustment was not assessed in this study, parent skills training plus BCT was superior to BCT and the control condition in reducing children's adjustment problems. Careful studies such as this are at the heart of translational science. They help move the field away from interventions that address a single problem in a narrowly defined population and toward the complexities that confront clinicians in the community. Initial evaluation research (albeit nonexperimental) has found Couples Coping Enhancement Training (CCET), a marital distress prevention program that includes content addressing stress and coping as well as communication and conflict skills, to be effective in improving couples' satisfaction and individual and dyadic coping with stress (e.g., Bodenmann, Pihet, Shantinath, Cina, & Widmer, 2006). To examine its effects in families with children, a sample of 100 couples with a child between the ages of 2 and 12 years, and who were “experiencing their children's upbringing to be demanding” were recruited. Families were assigned to either CCET or a no-treatment control condition, and it was hypothesized that CCET would reduce couples' conflicts about their children. Results indicated that CCET had positive effects on couples' communication and coping; however, effects on conflicts about children did not emerge (Ledermann, Bodenmann, & Cina, 2007).

Parenting Interventions that Reduce Interparental Conflict Interventions that focus on parenting have also been shown to exert indirect positive effects on parents' marital satisfaction, with those effects mediated by the improvements in the child's adjustment (McEachern et al., 2013). The Family Check-Up program is a parenting intervention designed to address emerging emotional and behavioral problems among toddlers

at high risk for those problems. The Family Check-Up program was founded on principles of motivational interviewing and takes a family-oriented approach to defining treatment targets and doses. Results of a randomized controlled trial with a high-risk sample of 731 families with a 2-year-old child indicated that children whose families participated in the Family Check-Up were less likely to have emotional, behavioral, or comorbid problems at age 4 than children in the control group, and the program's effects on child adjustment were mediated by changes in parenting (Connell et al., 2008; Dishion et al., 2008). A subsequent study of a subset of 435 of the original families indicated that reductions in child problems from age 2 to 4 predicted positive change in the parents' relationship satisfaction over the 3-year period from age 2 to 5 (McEachern et al., 2013). Similar findings emerged in a randomized controlled trial evaluating the effectiveness of the Oregon Parent Management Training intervention in improving marital functioning (Bullard et al., 2010). Over 30 years of rigorous empirical research supports the efficacy of Oregon Parent Management Training in addressing child conduct problems. This evaluation study included newly married couples in which the husband was the stepfather to the wife's biological children, and the target child for the study had five or more conduct problem symptoms. This parenting intervention had positive effects on marital functioning as well as on child adjustment problems, and these were maintained at the 2-year follow-up. Moreover, the effects on marital functioning were mediated by the improvements in children's behavior problems. For families in the no-treatment control condition, marital functioning, parenting practices, and child behavior problems all deteriorated over the follow-up period.

Mapping to the Organizational Framework and Future Directions Most of the interventions with a couple-oriented component are designed to alter the dynamics of interparental conflict, which may affect downstream changes in family functioning and can have effects on child functioning as well. The idea to offer particular kinds of prevention programs during key transition periods of family development seems wise (It has been advocated for close to two decades; e.g., Cowan & Cowan, 1995). Marriage education programs are already in wide use for couples planning to marry or newly married. Transitionto-parenthood programs that include a focus on interparental conflict are less common, and transition-to-school (elementary, middle, or high school) programs with interparental conflict and relevant parenting components are scarce outside the realm of empirical research. Such programs could capture families at points when family relational dynamics and the problems they confront are expected to change in important ways, and could provide couples with skills for managing conflict through the transition. In contrast to prevention programs designed to prevent problems before they emerge or become serious are interventions for families in which problems—whether couple problems, child problems, or both—are already identified. It appears from the brief review earlier that adding a BCT-type couple intervention to treatments that address other problems in the family (e.g., fathers' substance use; fathers' engagement with their children) or address parenting can reduce interparental conflict and improve the interparental relationship as well as child adjustment. At this juncture, although there are decades of research on couples therapy and

parent training, it is still unclear when and under what circumstances one of them might suffice to address both parent and child problems. Developmental considerations may also pertain to the selection and timing of intervention. For example, perhaps the timing of couples therapy has implications for its effects on child adjustment. If interparental conflict is addressed before it becomes entrenched or severe, or before it results in child adjustment problems, couples therapy may possibly be sufficient to prevent the onset or progression of child problems. Once the developmental unfolding of child adjustment problems reaches a certain point, however, a parenting intervention may be necessary to ameliorate them. Using the organizational framework as a guide, research that considers such factors can help advance theory and treatment research in this area.

Directions for Future Research Throughout this chapter, we have hinted at gaps in the empirical literature on interparental conflict and child adjustment as well as some of the important questions we believe researchers still need to address. The remainder of this chapter describes broad directions for scholars to consider in advancing knowledge in this area.

Improving Theoretical Integration and Precision Since the early 1990s, tremendous progress has been made in understanding how and why interparental conflict influences child adjustment. As noted earlier, much of this progress can be attributed to the development of theories specifically accounting for associations between interparental conflict and child adjustment. Investigators have begun to empirically test these theories and have identified a variety of psychological and biological processes that potentially explain associations between interparental conflict and child adjustment. One direction for future research is to examine more thoroughly how some of these hypothesized processes operate jointly in accounting for child adjustment or, even more precisely, in accounting for the effects of interparental conflict on child adjustment. Although there are some notable exceptions in the literature (e.g., Buehler et al., 2007; Davies et al., 2002), most investigators examine hypothesized mediators central to a single, specific theory. That is, investigators testing the cognitive contextual framework will focus on children appraisals of interparental conflict, whereas investigators testing EST focus on emotional insecurity of the interparental relationship. Of course, it makes sense to limit the focus when initially attempting to establish the efficacy of a theory. However, we believe we are now at a point at which it makes sense to begin to explore the integration of the various theoretical explanations. Investigating hypothesized mediators from different theories in tandem is one way to begin to do this; it seems plausible that several of the hypothesized mediating processes proposed by different theories simultaneously operate to influence child adjustment. Related to this point, it is important to develop and test theory that considers the complexity of how disparate processes (cognitive, emotional, behavioral, biological) might operate together to influence child adjustment within the context of destructive interparental conflict. A number

of plausible hypotheses can be generated and tested about how these different processes operate in tandem. For example, one plausible hypothesis is that multiple processes are important, and they exert additive or interactive effects on child adjustment. It is also plausible that a particular process explains the effects of most, or all, of the others. Furthermore, some processes may have stronger implications for longer-term problems than others. In addition, some processes may have stronger implications during certain developmental periods than others. Other, more complex hypotheses about how these processes operate in tandem can be proposed as well. It might also be fruitful to consider more thoroughly how destructive interparental conflict operates within the broader family context to influence children's adjustment. Investigators have already begun work in this area by examining how positive aspects of interparental functioning might contribute to positive child outcomes. Similarly, research is being conducted to examine how child functioning influences interparental conflict and how different family variables interact with destructive interparental conflict in the prediction of child problems. From a broader developmental psychopathology perspective, it will also be important for researchers to consider the transactions with and influences of the larger socio-ecological context in which children live and to gain a better understanding of environmental processes and events that contribute to children's risk and resilience in the face of interparental conflict. We believe it might also be advantageous to consider how destructive interparental conflict fits within other theories about the development of psychopathology. For example, many theories exist on the development of specific types of psychological disorders or the development of specific dimensions of psychopathology. It would be worthwhile for researchers to glean what is known from empirical tests of some of these theories and attempt to integrate concepts from these areas into research on interparental conflict and child adjustment.

Translating Research into Practice Interventions for children (and their families) exposed to destructive interparental conflict already exist. Similarly, interventions designed to reduce destructive interparental conflict are fairly widespread, especially within the marital/family therapy literatures. Although rigorous evaluations of interventions in this area are still rare, results from several well-designed studies show that these interventions can be effective in reducing child adjustment problems. It is important to note, however, that while some of these interventions are grounded in theory and research on interparental conflict and child adjustment, others are based only loosely on theory and research in this area. In addition, as with most clinical interventions and prevention programs, there is also considerable room for improvement. We believe an important direction for future research is to improve what we know about how to help children in families characterized by destructive interparental conflict. One way to do this is to understand more thoroughly how existing interventions work, especially those interventions that have been found to be effective in reducing child problems. For example, are these interventions altering specific processes (within the family and/or child) that might be leading to treatment gains? Are there key processes that need to be altered for

children/families to benefit from treatment? There is some evidence that changes in motherchild interaction contribute to treatment gains for children who have been exposed to frequent and severe IPV conflict (e.g., Jouriles et al., 2009), but much more work in this area is necessary. For example, processes (cognitive, emotional, behavioral, biological) that are hypothesized to be key in theories linking interparental conflict with child adjustment problems need to be examined more comprehensively in this intervention research. In addition, given that it is well established that destructive interparental conflict can lead to child adjustment problems and that effective interventions already exist, it should be a priority to ensure that the empirically supported interventions actually reach the children/families who can benefit from them. To date, most of the interventions reviewed in this chapter have remained largely in the confines of the academic settings in which they were developed. The dissemination and implementation of evidence-based interventions is a scientific field in its own right, but little has been done to formally disseminate interventions in this particular area of research.

Conclusion Substantial progress has been made in establishing a reliable link between interparental conflict and children's development as well as in developing and testing theory on how and why interparental conflict confers risk for child adjustment problems. In addition, progress is being made in translating this basic science into programs to promote higher-quality interparental relationships and optimal child development (in the context of interparental conflict). Yet many questions remain unanswered, especially for the development of a more nuanced understanding of the link between these two aspects of family life. Ultimately, a richer understanding of the processes underlying the relation between interparental conflict and child adjustment, and knowing which children are especially vulnerable (and resilient) to the effects of destructive interparental conflict, will allow for more effective intervention programs aimed at improving children's lives.

References Ablow, J. C., Measelle, J. R., Cowan, P. A., & Cowan, C. P. (2009). Linking marital conflict and children's adjustment: The role of young children's perceptions. Journal of Family Psychology, 23, 485–499. doi: 10.1037/a0015894 Ahmed, S., Koenig, M. A., & Stephenson, R. (2006). Effects of domestic violence on perinatal and early-childhood mortality: Evidence from North India. American Journal of Public Health, 96, 1423–1428. doi: 10.2105/AJPH.2005.066316 Amato, P. R., & McNeal, C. (1998). Parents' marital violence: Long-term consequences for children. Journal of Family Issues, 19, 123–139. doi: 10.1177/019251398019002001 Anderson, C. A., & Bushman, B. J. (2002). Human aggression. Annual Review of Psychology,

53, 27–51. doi: 10.1146/annurev.psych.53.100901.135231 Anderson, E. R., Hetherington, E., Reiss, D., & Howe, G. (1994). Parents' nonshared treatment of siblings and the development of social competence during adolescence. Journal of Family Psychology, 8, 303–320. doi: 10.1037/0893–3200.8.3.303 Appel, A. E., & Holden, G. W. (1998). The co-occurrence of spouse and physical child abuse: A review and appraisal. Journal of Family Psychology, 12, 578–599. doi: 10.1037/0893– 3200.12.4.578 Appleyard, K., Egeland, B., van Dulmen, M. M., & Sroufe, L. (2005). When more is not better: The role of cumulative risk in child behavior outcomes. Journal of Child Psychology and Psychiatry, 46, 235–245. doi: 10.1111/j.1469–7610.2004.00351.x Bair-Merritt, M. H., Jennings, J. M., Eaker, K., Tuman, J., Park, S., & Cheng, T. L. (2008). Screening for domestic violence and childhood exposure in families seeking care at an urban pediatric clinic. Journal of Pediatrics, 152, 734–736. doi: 10.1016/j.jpeds.2008.01.035 Bakermans-Kranenburg, M. J., & van IJzendoorn, M. H. (2008). Oxytocin receptor (OXTR) and serotonin transporter (5-HTT) genes associated with observed parenting. Social Cognitive and Affective Neuroscience, 3, 128–134. doi: 10.1093/scan/nsn004 Bandura, A. (1973). Aggression: A social learning analysis. Oxford, UK: Prentice-Hall. Bandura, A. (1986). Social foundations of thought and action: A social cognitive theory. Englewood Cliffs, NJ: Prentice-Hall. Bandura, A. (1994). Regulative function of perceived self-efficacy. In M. G. Rumsey, C. B. Walker, & J. Harris (Eds.), Personnel selection and classification (pp. 261–271). Hillsdale, NJ: Erlbaum. Basu, R., Malone, J. C., Levendosky, A. A., & Dubay, S (2009). Longitudinal treatment effectiveness outcomes of a group intervention for women and children exposed to domestic violence. Journal of Child & Adolescent Trauma, 2, 90–105. doi: 10.1080/19361520902880715 Beauchaine, T. P. (2001). Vagal tone, development, and Gray's motivational theory: Toward an integrated model of autonomic nervous system functioning in psychopathology. Development and Psychopathology, 13, 183–214. doi: 10.1017/S0954579401002012 Beauchaine, T. P. (2012). Physiological markers of emotion and behavior dysregulation in externalizing psychopathology. Monographs of the Society for Research in Child Development, 77, 79–86. doi: 10.1111/j.1540–5834.2011.00665.x Belsky, J. (1997). Attachment, mating, and parenting: An evolutionary interpretation. Human Nature, 8, 361–381. doi: 10.1007/BF02913039 Belsky, J., & Pluess, M. (2009). Beyond diathesis stress: Differential susceptibility to

environmental influences. Psychological Bulletin, 135, 885–908. doi: 10.1037/a0017376 Belsky, J., Youngblade, L., Rovine, M., & Volling, B. (1991). Patterns of marital change and parent–child interaction. Journal of Marriage and the Family, 53, 487–498. doi: 10.2307/352914 Berntson, G. G., & Cacioppo, J. T. (2004). Multilevel analyses and reductionism: Why social psychologists should care about neuroscience and vice versa. In J. T. Cacioppo & G. G. Berntson (Eds.), Essays in social neuroscience (pp. 107–120). Cambridge, MA: MIT Press. Bodenmann, G., Pihet, S., Shantinath, S. D., Cina, A., & Widmer, K. (2006). Improving dyadic coping in couples with a stress-oriented approach: A 2-year longitudinal study. Behavior Modification, 30, 571–597. doi: 10.1177/0145445504269902 Bornstein, M. H., & Suess, P. E. (2000). Physiological self-regulation and information processing in infancy: Cardiac vagal tone and habituation. Child Development, 71, 273–287. doi: 10.1111/1467–8624.00143 Boyce, W., & Ellis, B. J. (2005). Biological sensitivity to context: I. An evolutionarydevelopmental theory of the origins and functions of stress reactivity. Development and Psychopathology, 17, 271–301. doi: 10.1017/S0954579405050145 Bradford, K., Barber, B. K., Olsen, J. A., Maughan, S. L., Erickson, L. D., Ward, D., & Stolz, H. E. (2003). A multi-national study of interparental conflict, parenting, and adolescent functioning: South Africa, Bangladesh, China, India, Bosnia, Germany, Palestine, Colombia, and the United States. Marriage and Family Review, 35, 107–137. Buehler, C., Anthony, C., Krishnakumar, A., & Stone, G. (1997). Interparental conflict and youth problem behaviors: A meta-analysis. Journal of Child and Family Studies, 6, 223–247. doi: 10.1023/A:1025006909538 Buehler, C. Franck, K. L., & Cook, E. C. (2009). Adolescents' triangulation in marital conflict and peer relations. Journal of Research on Adolescence, 19, 669–689. doi: 10.1111/j.15327795.2009.00616.x Buehler, C., Lange, G., & Franck, K. L. (2007). Adolescents' cognitive and emotional responses to marital hostility. Child Development, 78, 775–789. doi: 10.1111/j.1467– 8624.2007.01032.x Bullard, L., Wachlarowicz, M., DeLeeuw, J., Snyder, J., Low, S., Forgatch, M., & DeGarmo, D. (2010). Effects of the Oregon model of Parent Management Training (PMTO) on marital adjustment in new stepfamilies: A randomized trial. Journal of Family Psychology, 24, 485– 96. doi: 10.1037/a0020267 Calkins, S. D., Graziano, P. A., & Keane, S. P. (2007). Cardiac vagal regulation differentiates among children at risk for behavior problems. Biological Psychology, 74, 144–153. doi: 10.1016/j.biopsycho.2006.09.005

Calkins, S. D., & Keane, S. P. (2004). Cardiac vagal regulation across the preschool period: Stability, continuity, and implications for childhood adjustment. Developmental Psychobiology, 45, 101–112. doi: 10.1002/dev.20020 Calvete, E., & Orue, I. (2011). The impact of violence exposure on aggressive behavior through social information processing in adolescents. American Journal of Orthopsychiatry, 81, 38–50. doi: 10.1111/j.1939–0025.2010.01070.x Chang, J., Theodore, A. D., Martin, S. L., & Runyan, D. K. (2008). Psychological abuse between parents: Associations with child maltreatment from a population-based sample. Child Abuse & Neglect, 32, 819–829. doi: 10.1016/j.chiabu.2007.11.003 Cicchetti, D. (1993). Developmental psychopathology: Reactions, reflections, projections. Developmental Review, 13, 471–502. doi: 10.1006/drev.1993.1021 Cicchetti, D., & Lynch, M. (1993). Toward an ecological/transactional model of community violence and child maltreatment: Consequences for children's development. Psychiatry: Interpersonal and Biological Processes, 56, 96–118. Clarke, S. B., Koenen, K. C., Taft, C. T., Street, A. E., King, L. A., & King, D. W. (2007). Intimate partner psychological aggression and child behavior problems. Journal of Traumatic Stress, 20, 97–101. doi: 10.1002/jts.20193 Cloitre, M., Stolbach, B. C., Herman, J. L., Pynoos, R. S., Wang, J., & Petkova, E. (2009). A developmental approach to complex PTSD: Childhood and adult cumulative trauma as predictors of symptom complexity. Journal of Traumatic Stress, 22, 399–408. doi: 10.1002/jts. Cohen, J. A., Mannarino, A. P., & Deblinger, E. (2006). Treating trauma and traumatic grief in children and adolescents. New York, NY: Guilford Press. Cohen, J. A., Mannarino, A. P., & Iyengar, S. (2011). Community treatment of posttraumatic stress disorder for children exposed to intimate partner violence: A randomized controlled trial. Archives of Pediatric and Adolescent Medicine, 165, 16–21. doi: 10.1001/archpediatrics.2010.247. Compas, B. E. (2009). Coping, regulation, and development during childhood and adolescence. New Directions for Child and Adolescent Development, 2009, 87–99. doi: 10.1002/cd.245 Compas, B. E., Connor-Smith, J. K., Saltzman, H., Thomsen, A., & Wadsworth, M. E. (2001). Coping with stress during childhood and adolescence: Problems, progress, and potential in theory and research. Psychological Bulletin, 127, 87–127. doi: 10.1037/0033–2909.127.1.87 Connell, A., Bullock, B. M., Dishion, T. J., Shaw, D., Wilson, M., & Gardner, F. (2008). Family intervention effects on co-occurring early childhood behavioral and emotional problems: A latent transition analysis approach. Journal of Abnormal Child Psychology, 36,

1211–25. doi: 10.1007/s10802–008–9244–6 Cowan, C. P., Cowan, P. A., & Barry, J. (2011). Couples' groups for parents of preschoolers: Ten-year outcomes of a randomized trial. Journal of Family Psychology, 25, 240–50. doi: 10.1037/a0023003 Cowan, C. P., & Cowan, P. A. (1995). Interventions to ease the transition to parenthood: Why they are needed and what they can do. Family Relations, 44, 412–423. Cowan, P. A., Cowan, C. P., Cohen, N., Pruett, M. K., & Pruett, K. (2008). Supporting fathers' engagement with their kids. In J. D. Berrick & N. Gilbert (Eds.), Raising children: Emerging needs, modern risks, and social responses (pp. 44–80). New York, NY: Oxford University Press. Cowan, P. A., Cowan, C. P., Pruett, M., Pruett, K., & Wong, J. J. (2009). Promoting fathers' engagement with children: Preventive interventions for low-income families. Journal of Marriage and Family, 71, 663–679. doi: 10.1111/j.1741–3737.2009.00625.x Cox, M. J., Paley, B., & Harter, K. (2001). Interparental conflict and parent–child relationships. In J. H. Grych & F. D. Fincham (Eds.), Interparental conflict and child development: Theory, research, and applications (pp. 249–272). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511527838.011 Crick, N. R., & Dodge, K. A. (1994). A review and reformulation of social informationprocessing mechanisms in children's social adjustment. Psychological Bulletin, 115, 74–101. doi: 10.1037/0033–2909.115.1.74 Criss, M. M., Pettit, G. S., Bates, J. E., Dodge, K. A., & Lapp, A. L. (2002). Family adversity, positive peer relationships, and children's externalizing behavior: A longitudinal perspective on risk and resilience. Child Development, 73, 1220–1237. doi: 10.1111/1467–8624.00468 Crockenberg, S., & Langrock, A. (2001). The role of specific emotions in children's responses to interparental conflict: A test of the model. Journal of Family Psychology, 15, 163–182. doi: 10.1037/0893–3200.15.2.163 Cui, M., Conger, R. D., & Lorenz, F. O. (2005). Predicting change in adolescent adjustment from change in marital problems. Developmental Psychology, 41, 812–823. doi: 10.1037/0012–1649.41.5.812 Cui, M., Durtschi, J. A., Donnellan, M., Lorenz, F. O., & Conger, R. D. (2010). Intergenerational transmission of relationship aggression: A prospective longitudinal study. Journal of Family Psychology, 24, 688–697. doi: 10.1037/a0021675 Cummings, E. M., Ballard, M., El-Sheikh, M., & Lake, M. (1991). Resolution and children's responses to interadult anger. Developmental Psychology, 27, 462–470. doi: 10.1037/0012– 1649.27.3.462

Cummings, E. M., & Cummings, J. L. (1988). A process-oriented approach to children's coping with adults' angry behavior. Developmental Review, 8, 296–321. doi: 10.1016/0273– 2297(88)90008–1 Cummings, E. M., & Davies, P. T. (2010). Marital conflict and children: An emotional security perspective. New York, NY: Guilford Press. Cummings, E. M., Faircloth, W. B., Mitchell, P. M., Cummings, J. S., & Schermerhorn, A. C. (2008). Evaluating a brief prevention program for improving marital conflict in community families. Journal of Family Psychology, 22, 193–202. doi: 10.1037/0893–3200.22.2.193 Cummings, E. M., Goeke-Morey, M. C., & Papp, L. M. (2004). Everyday marital conflict and child aggression. Journal of Abnormal Child Psychology, 32, 191–202. doi: 10.1023/B:JACP.0000019770.13216.be Cummings, E. M., Iannotti, R. J., & Zahn-Waxler, C. (1985). Influence of conflict between adults on the emotions and aggression of young children. Developmental Psychology, 21, 495– 507. doi: 10.1037/0012–1649.21.3.495 Cummings, E. M., Schermerhorn, A. C., Davies, P. T., Goeke-Morey, M. C., & Cummings, J. S. (2006). Interparental discord and child adjustment: Prospective investigations of emotional security as an explanatory mechanism. Child Development, 77, 132–152. doi: 10.1111/j.1467– 8624.2006.00861.x Cummings, E. M., Vogel, D., Cummings, J. S., & El-Sheikh, M. (1989). Children's responses to different forms of expression of anger between adults. Child Development, 60, 1392–1404. doi: 10.2307/1130929 Cummings, E. M., Zahn-Waxler, C., & Radke-Yarrow, M. (1981). Young children's responses to expressions of anger and affection by others in the family. Child Development, 52, 1274– 1282. doi: 10.2307/1129516 Dadds, M. R., Atkinson, E., Turner, C., Blums, G., & Lendich, B. (1999). Family conflict and child adjustment: Evidence for a cognitive-contextual model of intergenerational transmission. Journal of Family Psychology, 13, 194–208. doi: 10.1037/0893–3200.13.2.194 Davies, P. T., & Cicchetti, D. (2004). Toward an integration of family systems and developmental psychopathology approaches. Development and Psychopathology, 16, 477– 481. doi: 10.1017/S0954579404004626 Davies, P. T., & Cummings, E. M. (1994). Marital conflict and child adjustment: An emotional security hypothesis. Psychological Bulletin, 116, 387–411. doi: 10.1037/0033– 2909.116.3.387 Davies, P. T., & Cummings, E. M. (1998). Exploring children's emotional security as a mediator of the link between marital relations and child adjustment. Child Development, 69, 124–139. doi: 10.2307/1132075

Davies, P. T., & Cummings, E. (2006). Interparental discord, family process, and developmental psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, Vol. 3: Risk, disorder, and adaptation (2nd ed.) (pp. 86–128). Hoboken, NJ: Wiley. Davies, P. T., & Forman, E. M. (2002). Children's patterns of preserving emotional security in the interparental subsystem. Child Development, 73, 1880–1903. doi: 10.1111/1467– 8624.t01–1–00512 Davies, P. T., Forman, E. M., Rasi, J. A., & Stevens, K. I. (2002). Assessing children's emotional security in the interparental relationship: The Security in the Interparental Subsystem Scales. Child Development, 73, 544–562. doi: 10.1111/1467–8624.00423 Davies, P. T., Harold, G. T., Goeke-Morey, M. C., & Cummings, E. (2002). Child emotional security and interparental conflict. Monographs of the Society for Research in Child Development, 67, vii–viii. doi: 10.1111/1540–5834.00205 Davies, P. T., Manning, L. G., & Cicchetti, D. (2013). Tracing the cascade of children's insecurity in the interparental relationship: The role of stage-salient tasks. Child Development, 84, 297–312. doi: 10.1111/j.1467–8624.2012.01844.x Davies, P. T., Martin, M. J., & Cicchetti, D. (2012). Delineating the sequelae of destructive and constructive interparental conflict for children within an evolutionary framework. Developmental Psychology, 48, 939–955. doi: 10.1037/a0025899 Davies, P. T., Myers, R. L., & Cummings, E. (1996). Responses of children and adolescents to marital conflict scenarios as a function of the emotionality of conflict endings. Merrill-Palmer Quarterly, 42, 1–21. Davies, P. T., & Sturge-Apple, M. (2007). Advances in the formulation of emotional security theory: An ethologically based perspective. In R. V. Kail (Ed.), Advances in child development and behavior (Vol. 35, pp. 87–137). San Diego, CA: Elsevier Academic Press. Davies, P. T., Sturge-Apple, M. L., & Cicchetti, D. (2011). Interparental aggression and children's adrenocortical reactivity: Testing an evolutionary model of allostatic load. Development and Psychopathology, 23, 801–814. doi: 10.1017/S0954579411000319 Davies, P. T., Sturge-Apple, M. L., Cicchetti, D., & Cummings, E. M. (2007). The role of child adrenocortical functioning in pathways between interparental conflict and child maladjustment. Developmental Psychology, 43, 918–930. doi: 10.1037/0012–1649.43.4.918 Davies, P. T., Sturge-Apple, M. L., Cicchetti, D., Manning, L. G., & Zale, E. (2009). Children's patterns of emotional reactivity to conflict as explanatory mechanisms in links between interpartner aggression and child physiological functioning. Journal of Child Psychology and Psychiatry, 50, 1384–1391. doi: 10.1111/j.1469–7610.2009.02154.x Davies, P. T., Sturge-Apple, M. L., Woitach, M. J., & Cummings, E. (2009). A process analysis

of the transmission of distress from interparental conflict to parenting: Adult relationship security as an explanatory mechanism. Developmental Psychology, 45, 1761–1773. doi: 10.1037/a0016426 Davis, B. T., Hops, H., Alpert, A., & Sheeber, L. (1998). Child responses to parental conflict and their effect on adjustment: A study of triadic relations. Journal of Family Psychology, 12, 163–177. doi: 10.1037/0893–3200.12.2.163 De Arellano, M. A., Lyman, D. R., Jobe-Shields, L., George, P., Dougherty, R. H., Daniels, A. S., …Delphin-Rittmon, M. E. (2014). Trauma-focused cognitive-behavioral therapy for children and adolescents: Assessing the evidence. Psychiatric Services, 65, 591–602. doi: 10.1176/appi.ps.201300255 De Bellis, M. (1999). Developmental traumatology part I: Biological stress systems. Biological Psychiatry, 45, 1259–1270. doi: 10.1016/S0006–3223(99)00044-X Deblinger, E., Mannarino, A. P., Cohen, J. A., Runyon, M. K., & Steer, R. A. (2011). Traumafocused cognitive behavioral therapy for children: Impact of the trauma narrative and treatment length. Depression and Anxiety, 28, 67–75. doi: 10.1002/da.20744 DeBoard-Lucas, R. L., Fosco, G. M., Raynor, S. R., & Grych, J. H. (2010). Interparental conflict in context: Exploring relations between parenting processes and children's conflict appraisals. Journal of Clinical Child and Adolescent Psychology, 39, 163–175. Del Giudice, M., Ellis, B. J., & Shirtcliff, E. A. (2011). The Adaptive Calibration Model of stress responsivity. Neuroscience and Biobehavioral Reviews, 35, 1562–1592. doi: 10.1016/j.neubiorev.2010.11.007 Del Giudice, M., Hinnant, J., Ellis, B. J., & El-Sheikh, M. (2012). Adaptive patterns of stress responsivity: A preliminary investigation. Developmental Psychology, 48, 775–790. doi: 10.1037/a0026519 Dietrich, A., Riese, H., Sondeijker, F. L., Greaves-Lord, K., van Roon, A. M., Ormel, J., … Rosmalen, J. M. (2007). Externalizing and internalizing problems in relation to autonomic function: A population-based study in preadolescents. Journal of the American Academy of Child & Adolescent Psychiatry, 46, 378–386. doi: 10.1097/CHI.0b013e31802b91ea Dishion, T. J., Shaw, D., Connell, A., Gardner, F., Weaver, C., & Wilson, M. (2008). The family check-up with high-risk indigent families: Preventing problem behavior by increasing parents' positive behavior support in early childhood. Child Development, 79, 1395–1414. doi: 10.1111/j.1467–8624.2008.01195.x Dodge, K. A. (1986). A social information processing model of social competence in children. In M. Perlmutter (Ed.), The Minnesota symposium on child psychology (Vol. 18, pp. 77–125). Hillsdale, NJ: Erlbaum. Du Rocher Schudlich, T. D., White, C. R., Fleischhauer, E. A., & Fitzgerald, K. A. (2011).

Observed infant reactions during live interparental conflict. Journal of Marriage and Family, 73, 221–235. doi: 10.1111/j.1741–3737.2010.00800.x Ehrensaft, M. K., Cohen, P., Brown, J., Smailes, E., Chen, H., & Johnson, J. G. (2003). Intergenerational transmission of partner violence: A 20-year prospective study. Journal of Consulting and Clinical Psychology, 71, 741–753. doi: 10.1037/0022–006X.71.4.741 Ellis, B. J., Boyce, W., Belsky, J., Bakermans-Kranenburg, M. J., & Van Ijzendoorn, M. H. (2011). Differential susceptibility to the environment: An evolutionary–neurodevelopmental theory. Development and Psychopathology, 23, 7–28. doi: 10.1017/S0954579410000611 Ellis, B. J., Essex, M. J., & Boyce, W. (2005). Biological sensitivity to context: II. Empirical explorations of an evolutionary-developmental theory. Development and Psychopathology, 17, 303–328. doi: 10.1017/S0954579405050157 El-Sheikh, M. (2005). The role of emotional responses and physiological reactivity in the marital conflict-child functioning link. Journal of Child Psychology and Psychiatry, 46, 1191–1199. doi: 10.1111/j.1469–7610.2005.00418.x El-Sheikh, M. (2007). Children's skin conductance level and reactivity: Are these measures stable over time and across tasks? Developmental Psychobiology, 49, 180–186. doi: 10.1002/dev.20171 El-Sheikh, M., Arsiwalla, D. D., Hinnant, J., & Erath, S. A. (2011). Children's internalizing symptoms: The role of interactions between cortisol and respiratory sinus arrhythmia. Physiology & Behavior, 103, 225–232. doi: 10.1016/j.physbeh.2011.02.004 El-Sheikh, M., Buckhalt, J. A., Cummings, E. M., & Keller, P. (2007). Sleep disruptions and emotional insecurity are pathways of risk for children. Journal of Child Psychology and Psychiatry, 48, 88–96. doi: 10.1111/j.1469–7610.2006.01604.x El-Sheikh, M., Cummings, E. M., Kouros, C. D., Elmore-Staton, L., & Buckhalt, J. (2008). Marital psychological and physical aggression and children's mental and physical health: Direct, mediated, and moderated effects. Journal of Consulting and Clinical Psychology, 76, 138–148. doi: 10.1037/0022–006X.76.1.138 El-Sheikh, M., Cummings, E. M., & Reiter, S. (1996). Preschoolers' responses to ongoing interadult conflict: The role of prior exposure to resolved versus unresolved arguments. Journal of Abnormal Child Psychology, 24, 665–679. doi: 10.1007/BF01670106 El-Sheikh, M., & Erath, S. A. (2011). Family conflict, autonomic nervous system functioning, and child adaptation: State of the science and future directions. Development and Psychopathology, 23, 703–721. doi: 10.1017/S0954579411000034 El-Sheikh, M., Harger, J., & Whitson, S. M. (2001). Exposure to interparental conflict and children's adjustment and physical health: The moderating role of vagal tone. Child Development, 72, 1617–1636. doi: 10.1111/1467–8624.00369

El-Sheikh, M., Keiley, M., Erath, S., & Dyer, W. (2013). Marital conflict and growth in children's internalizing symptoms: The role of autonomic nervous system activity. Developmental Psychology, 49, 92–108. doi: 10.1037/a0027703 El-Sheikh, M., Keller, P. S., & Erath, S. A. (2007). Marital conflict and risk for child maladjustment over time: Skin conductance level reactivity as a vulnerability factor. Journal of Abnormal Child Psychology, 35, 715–727. doi: 10.1007/s10802–007–9127–2 El-Sheikh, M., Kouros, C. D., Erath, S., Cummings, E. M., Keller, P. S., & Elmore-Staton, L. (2009). Marital conflict and children's externalizing behavior: Interactions between parasympathetic and sympathetic nervous system activity. Monographs of the Society for Research in Child Development, 74, 1–101. El-Sheikh, M., & Whitson, S. A. (2006). Longitudinal relations between marital conflict and child adjustment: Vagal regulation as a protective factor. Journal of Family Psychology, 20, 30–39. doi: 10.1037/0893–3200.20.1.30 Emery, C. R. (2009). Stay for the children? Husband violence, marital stability, and children's behavior problems. Journal of Marriage and Family, 71, 905–916. doi: 10.1111/j.1741– 3737.2009.00643.x Emery, R. E. (1982). Interparental conflict and the children of discord and divorce. Psychological Bulletin, 92, 310–330. doi: 10.1037/0033–2909.92.2.310 Emery, R. E. (1989). Family violence. American Psychologist, 44, 321–328. doi: 10.1037/0003–066X.44.2.321 Emery, R. E., & Kitzmann, K. M. (1995). The child in the family: Disruptions in family functions. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, Vol. 2: Risk, disorder, and adaptation (pp. 3–31). Oxford, UK: Wiley. Erel, O., & Burman, B. (1995). Interrelatedness of marital relations and parent-child relations: A meta-analytic review. Psychological Bulletin, 118, 108–132. doi: 10.1037/0033– 2909.118.1.108 Erel, O., Margolin, G., & John, R. S. (1998). Observed sibling interaction: Links with the marital and the mother–child relationship. Developmental Psychology, 34, 288–298. doi: 10.1037/0012–1649.34.2.288 Evans, S. E., Davies, C., & DiLillo, D. (2008). Exposure to domestic violence: A metaanalysis of child and adolescent outcomes. Aggression and Violent Behavior, 13, 131–140. doi: 10.1016/j.avb.2008.02.005 Faircloth, W. B., Schermerhorn, A. C., Mitchell, P. M., Cummings, J. S., & Cummings, E. M. (2011). Testing the long-term efficacy of a prevention program for improving marital conflict in community families. Journal of Applied Developmental Psychology, 32, 189–197. doi: 10.1016/j.appdev.2011.05.004

Fantuzzo, J. W., & Fusco, R. A. (2007). Children's direct exposure to types of domestic violence crime: A population-based investigation. Journal of Family Violence, 22, 543–552. doi: 10.1007/s10896–007–9105-z Fauber, R., Forehand, R., Thomas, A. M., & Wierson, M. (1990). A mediational model of the impact of marital conflict on adolescent adjustment in intact and divorced families: The role of disrupted parenting. Child Development, 61, 1112–1123. doi: 10.2307/1130879 Fehringer, J. A., & Hindin, M. J. (2009). Like parent, like child: Intergenerational transmission of partner violence in Cebu, the Philippines. Journal of Adolescent Health, 44, 363–371. doi: 10.1016/j.jadohealth.2008.08.012 Feinberg, M. E., Jones, D. E., Kan, M. L., & Goslin, M. C. (2010). Effects of family foundations on parents and children: 3.5 years after baseline. Journal of Family Psychology, 24, 532–42. doi: 10.1037/a0020837 Fergusson, D. M., Boden, J. M., & Horwood, L. (2006). Examining the intergenerational transmission of violence in a New Zealand birth cohort. Child Abuse & Neglect, 30, 89–108. doi: 10.1016/j.chiabu.2005.10.006 Finkelhor, D., Ormrod, R. K., & Turner, H. A. (2007). Poly-victimization: A neglected component in child victimization. Child Abuse & Neglect, 31, 7–26. doi: 10.1016/j.chiabu.2006.06.008 Fite, J. E., Bates, J. E., Holtzworth-Munroe, A., Dodge, K. A., Nay, S., Pettit, G. S. (2008). Social information processing mediates the intergenerational transmission of aggressiveness in romantic relationships. Journal of Family Psychology, 22, 367–376. Fixsen, D. L., Naoom, S. F., Blase, K. A., Friedman, R. M., & Wallace, F. (2005). Implementation research: A synthesis of the literature (pp. 1–125). Tampa, FL: University of South Florida, Louis de la Parta Florida Mental Health Institute, The National Implementation Research Network. (FMHI Publication #231). Flach, C., Leese, M., Heron, J., Evans, J., Feder, G., Sharp, D., & Howard, L. (2011). Antenatal domestic violence, maternal mental health and subsequent child behaviour: A cohort study. BJOG: An International Journal of Obstetrics & Gynaecology, 118(11), 1383–1391. doi: 10.1111/j.1471–0528.2011.03040.x Fosco, G. M., DeBoard, R. L., & Grych, J. H. (2007). Making sense of family violence: Implications of children's appraisals of interparental aggression for their short- and long-term functioning. European Psychologist, 12, 6–16. doi: 10.1027/1016–9040.12.1.6 Fosco, G. M., & Feinberg, M. E. (2015). Cascading effects of interparental conflict in adolescence: Linking threat appraisals, self-efficacy, and adjustment. Development and Psychopathology, 27, 239–252. Fosco, G. M., & Grych, J. H. (2007). Emotional expression in the family as a context for

children's appraisals of interparental conflict. Journal of Family Psychology, 21, 248–258. doi: 10.1037/0893–3200.21.2.248 Fosco, G. M., & Grych, J. H. (2008). Emotional, cognitive, and family systems mediators of children's adjustment to interparental conflict. Journal of Family Psychology, 22, 843–854. doi: 10.1037/a0013809 Fosco, G. M., & Grych, J. H. (2010). Adolescent triangulation into parental conflicts: Longitudinal implications for appraisals and adolescent-parent relations. Journal of Marriage and Family, 72, 254–266. doi: 10.1111/j.1741-3737.2010.00697.x Framo, J. L. (1965). Rationale and techniques of intensive family therapy. In I. BoszormenyiNagy & J. Framo (Eds.), Intensive family therapy: Theoretical and practical aspects (pp. 143–212). New York, NY: Harper & Row. Frosch, C. A., & Mangelsdorf, S. C. (2001). Marital behavior, parenting behavior, and multiple reports of preschoolers' behavior problems: Mediation or moderation? Developmental Psychology, 37, 502–519. doi: 10.1037/0012–1649.37.4.502 Garmezy, N., Masten, A. S., & Tellegen, A. (1984). The study of stress and competence in children: A building block for developmental psychopathology. Child Development, 55, 97– 111. doi: 10.2307/1129837 Gattis, K. S., Simpson, L. E., & Christensen, A. (2008). What about the kids? Parenting and child adjustment in the context of couple therapy. Journal of Family Psychology, 22, 833–42. doi: 10.1037/a0013713 Gerard, J. M., Buehler, C., Franck, K., & Anderson, O. (2005). In the eyes of the beholder: Cognitive appraisals as mediators of the association between interparental conflict and youth maladjustment. Journal of Family Psychology, 19, 376–384. doi: 10.1037/0893– 3200.19.3.376 Gerard, J. M., Krishnakumar, A., & Buehler, C. (2006). Marital conflict, parent-child relations, and youth maladjustment: A longitudinal investigation of spillover effects. Journal of Family Issues, 27, 951–975. doi: 10.1177/0192513X05286020 Ghazarian, S., & Buehler, C. (2010). Interparental conflict and academic achievement: An examination of mediating and moderating factors. Journal of Youth and Adolescence, 39, 23– 32. Goeke-Morey, M. C., Cummings, E., Harold, G. T., & Shelton, K. H. (2003). Categories and continua of destructive and constructive marital conflict tactics from the perspective of U.S. and Welsh children. Journal of Family Psychology, 17, 327–338. doi: 10.1037/0893– 3200.17.3.327 Goeke-Morey, M. C., Cummings, E., & Papp, L. M. (2007). Children and marital conflict resolution: Implications for emotional security and adjustment. Journal of Family Psychology,

21, 744–753. doi: 10.1037/0893–3200.21.4.744 Gomulak-Cavicchio, B. M., Davies, P. T., & Cummings, E. (2006). The role of maternal communication patterns about interparental disputes in associations between interparental conflict and child psychological maladjustment. Journal of Abnormal Child Psychology, 34, 757–771. doi: 10.1007/s10802–006–9050-y Gonzales, N. A., Pitts, S. C., Hill, N. E., & Roosa, M. W. (2000). A mediational model of the impact of interparental conflict on child adjustment in a multiethnic, low-income sample. Journal of Family Psychology, 14, 365–379. doi: 10.1037/0893–3200.14.3.365 Gottman, J. M., & Katz, L. F. (1989). Effects of marital discord on young children's peer interaction and health. Developmental Psychology, 25, 373–381. doi: 10.1037/0012– 1649.25.3.373 Graham, A. M., Kim, H. K., & Fisher, P. A. (2012). Partner aggression in high-risk families from birth to age 3 years: Associations with harsh parenting and child maladjustment. Journal of Family Psychology, 26, 105–114. doi: 10.1037/a0026722 Graham-Bermann, S. A. (2000). Evaluating interventions for children exposed to family violence. Journal of Aggression, Maltreatment & Trauma, 4, 191–215. Graham-Bermann, S. A. (1992). The Kids' Club: A preventive intervention program for children of battered women. Department of Psychology, University of Michigan. Graham-Bermann, S. A., & Halabu, H. M. (2004). Fostering resilient coping in children exposed to violence: Cultural considerations. In P. G. Jaffe, L. L. Baker, & A. J. Cunningham (Eds.), Protecting children from domestic violence: Strategies for community intervention (pp. 71–88). New York, NY: Guilford Press. Graham-Bermann, S. A., & Hughes, H. M. (2003). Intervention for children exposed to interparental violence (IPV): Assessment of needs and research priorities. Clinical Child and Family Psychology Review, 6, 189–204. doi: 10.1023/A:1024962400234 Graham-Bermann, S. A., & Levendosky, A. A. (1998). Traumatic stress symptoms in children of battered women. Journal of Interpersonal Violence, 14, 111–128. Graham-Bermann, S. A., Lynch, S., Banyard, V. L., Devoe, E. R., & Halabu, H. M. (2007). Community-based intervention for children exposed to intimate partner violence: An efficacy trial. Journal of Consulting and Clinical Psychology, 75, 199–209. doi: 10.1037/0022– 006X.75.2.199 Grych, J. H. (1998). Children's appraisals of interparental conflict: Situational and contextual influences. Journal of Family Psychology, 12, 437–453. doi: 10.1037/0893–3200.12.3.437 Grych, J. H., & Cardoza-Fernandes, S. (2001). Understanding the impact of interparental conflict on children: The role of social cognitive processes. In J. H. Grych & F. D. Fincham

(Eds.), Interparental conflict and child development: Theory, research, and applications (pp. 157–187). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511527838.008 Grych, J. H., & Fincham, F. D. (1990). Marital conflict and children's adjustment: A cognitivecontextual framework. Psychological Bulletin, 108, 267–290. doi: 10.1037/0033– 2909.108.2.267 Grych, J. H., & Fincham, F. D. (1993). Children's appraisals of marital conflict: Initial investigations of the cognitive-contextual framework. Child Development, 64, 215–230. doi: 10.2307/1131447 Grych, J. H., Fincham, F. D., Jouriles, E. N., & McDonald, R. (2000). Interparental conflict and child adjustment: Testing the mediational role of appraisals in the cognitive-contextual framework. Child Development, 71, 1648–1661. doi: 10.1111/1467–8624.00255 Grych, J. H., Harold, G. T., & Miles, C. J. (2003). A prospective investigation of appraisals as mediators of the link between interparental conflict and child adjustment. Child Development, 74, 1176–1193. doi: 10.1111/1467–8624.00600 Grych, J. H., Jouriles, E. N., Swank, P. R., McDonald, R., & Norwood, W. D. (2000). Patterns of adjustment among children of battered women. Journal of Consulting and Clinical Psychology, 68, 84–94. doi: 10.1037/0022–006X.68.1.84 Grych, J. H., Seid, M., & Fincham, F. (1992). Assessing marital conflict from the child's perspective: the Children's Perception of Interparental Conflict Scale. Child Development, 63, 558–572. Gunnar, M., & Quevedo, K. (2007). The neurobiology of stress and development. Annual Review of Psychology, 58, 145–173. Hamby, S., Finkelhor, D., Turner, H., & Ormrod, R. (2011). Children's exposure to intimate partner violence and other family violence (NCJ232272). Washington, DC: US Department of Justice. Harold, G. T., Aitken, J. J., & Shelton, K. H. (2007). Inter-parental conflict and children's academic attainment: a longitudinal analysis. Journal of Child Psychology & Psychiatry, 48(12), 1223–1232. Harold, G. T., & Conger, R. (1997). Marital conflict and adolescent distress: The role of adolescent awareness. Child Development, 68, 333–350. Harold, G. T., Elam, K. K., Lewis, G., Rice, F., & Thapar, A. (2012). Interparental conflict, parent psychopathology, hostile parenting, and child antisocial behavior: Examining the role of maternal versus paternal influences using a novel genetically sensitive research design. Development and Psychopathology, 24, 1283–1295. doi: 10.1017/S0954579412000703

Harold, G. T., Fincham, F. D., Osborne, L. N., & Conger, R. D. (1997). Mom and dad are at it again: Adolescent perceptions of marital conflict and adolescent psychological distress. Developmental Psychology, 33, 333–350. Harold, G. T., Leve, L. D., Elam, K. K., Thapar, A., Neiderhiser, J. M., Natsuaki, M. N., … Reiss, D. (2013). The nature of nurture: Disentangling passive genotype–environment correlation from family relationship influences on children's externalizing problems. Journal of Family Psychology, 27, 12–21. doi: 10.1037/a0031190 Hawkins, A. J., Blanchard, V. L., Baldwin, S. A., & Fawcett, E. B. (2008). Does marriage and relationship education work? A meta-analytic study. Journal of Consulting and Clinical Psychology, 76, 723–34. doi: 10.1037/a0012584 Hinnant, J., Elmore-Staton, L., & El-Sheikh, M. (2011). Developmental trajectories of respiratory sinus arrythmia and preejection period in middle childhood. Developmental Psychobiology, 53, 59–68. doi: 10.1002/dev.20487 Hinnant, J., & El-Sheikh, M. (2009). Children's externalizing and internalizing symptoms over time: The role of individual differences in patterns of RSA responding. Journal of Abnormal Child Psychology, 37, 1049–1061. doi: 10.1007/s10802–009–9341–1 Hubbard, R. M., & Adams, C. F. (1936). Factors affecting the success of child guidance clinic treatment. American Journal of Orthopsychiatry, 6, 81–102. Huth-Bocks, A. C., Levendosky, A. A., Bogat, G., & von Eye, A. (2004). The impact of maternal characteristics and contextual variables on infant-mother attachment. Child Development, 75, 480–496. doi: 10.1111/j.1467–8624.2004.00688.x Ireland, T. O., & Smith, C. A. (2009). Living in partner-violent families: Developmental links to antisocial behavior and relationship violence. Journal of Youth and Adolescence, 38, 323– 339. doi: 10.1007/s10964–008–9347-y Jaffe, P., Wolfe, D. A., & Campbell, M. (2012). Growing up with domestic violence: Assessment, intervention, and prevention strategies for children and adolescents. Cambridge, MA: Hogrefe. Jaffee, S. R., Moffitt, T. E., Caspi, A., & Taylor, A. (2003). Life with (or without) father: The benefits of living with two biological parents depend on the father's antisocial behavior. Child Development, 74, 109–126. doi: 10.1111/1467–8624.t01–1–00524 Jaffee, S. R., & Poulton, R. (2006). Reciprocal effects of mothers' depression and children's problem behaviors from middle childhood to early adolescence. In A. C. Huston & M. N. Ripke (Eds.), Developmental contexts in middle childhood: Bridges to adolescence and adulthood (pp. 107–129). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511499760.007 Jenkins, J. M. (2000). Marital conflict and children's emotions: The development of an anger

schema. Journal of Marriage and the Family, 62, 723–736. Jenkins, J. M., Smith, M. A., & Graham, P. J. (1989). Coping with parental quarrels. Journal of the American Academy of Child & Adolescent Psychiatry, 28, 182–189. doi: 10.1097/00004583–198903000–00006 Johnston, J. R., Gonzàlez, R., & Campbell, L. E. (1987). Ongoing postdivorce conflict and child disturbance. Journal of Abnormal Child Psychology, 15, 493–509. doi: 10.1007/BF00917236 Jouriles, E. N., Bourg,W. J., & Farris, A. M. (1991). Marital adjustment and child conduct problems: A comparison of the correlation across subsamples. Journal of Consulting and Clinical Psychology, 59, 354–357. Jouriles, E. N., Brown, A. S., McDonald, R., Rosenfield, D., Leahy, M. M., & Silver, C. (2008). Intimate partner violence and preschoolers' explicit memory functioning. Journal of Family Psychology, 22, 420–428. Jouriles, E. N., & Farris, A. M. (1992). Effects of marital conflict on subsequent parent-son interactions. Behavior Therapy, 23, 355–374. doi: 10.1016/S0005–7894(05)80163–7 Jouriles, E. N., McDonald, R., Mueller, V., & Grych, J. H. (2012). Youth experiences of family violence and teen dating violence perpetration: Cognitive and emotional mediators. Clinical Child and Family Psychology Review, 15, 58–68. doi: 10.1007/s10567–011–0102–7 Jouriles, E. N., McDonald, R., Norwood, W. D., Ware, H., Spiller, L., & Swank, P. R. (1998). Knives, guns, and interparent violence: Relations with child behavior problems. Journal of Family Psychology, 12, 178–194. doi: 10.1037/0893–3200.12.2.178 Jouriles, E. N., McDonald, R., Rosenfield, D., Stephens, N., Corbitt-Shindler, D., & Miller, P. C. (2009). Reducing conduct problems among children exposed to intimate partner violence: A randomized clinical trial examining effects of Project Support. Journal of Consulting and Clinical Psychology, 77, 705–717. doi: 10.1037/a0015994 Jouriles, E. N., McDonald, R., Slep, A., Heyman, R. E., & Garrido, E. (2008). Child abuse in the context of domestic violence: Prevalence, explanations, and practice implications. Violence and Victims, 23, 221–235. doi: 10.1891/0886–6708.23.2.221 Jouriles, E. N., McDonald, R., Spiller, L. C., Norwood, W. D., Swank, P. R., Stephens, N., … Buzy, W. M. (2001). Reducing conduct problems among children of battered women. Journal of Consulting and Clinical Psychology, 69, 774–785. doi: 10.1037//0022–006X.69.5.774 Jouriles, E. N., Mueller, V., Rosenfield, D., McDonald, R., & Dodson, M. (2012). Teens' experiences of harsh parenting and exposure to severe intimate partner violence: Adding insult to injury in predicting teen dating violence. Psychology of Violence, 2, 125–138. doi: 10.1037/a0027264

Jouriles, E. N., Murphy, C. M., Farris, A. M., Smith, D. A., Richters, J. E., & Waters, E. (1991). Marital adjustment, parental disagreements about child rearing, and behavior problems in boys: Increasing the specificity of the marital assessment. Child Development, 62, 1424– 1433. doi: 10.2307/1130816 Jouriles, E. N., Murphy, C. M., & O'Leary, K. (1989). Interspousal aggression, marital discord, and child problems. Journal of Consulting and Clinical Psychology, 57, 453–455. doi: 10.1037/0022–006X.57.3.453 Jouriles, E. N., Norwood, W. D., McDonald, R., Vincent, J. P., & Mahoney, A. (1996). Physical violence and other forms of marital aggression: Links with children's behavior problems. Journal of Family Psychology, 10, 223–234. doi: 10.1037/0893–3200.10.2.223 Jouriles, E. N., Spiller, L., Stephens, N., McDonald, R., & Swank, P. (2000). Variability in adjustment of children of battered women: The role of child appraisals of interparent conflict. Cognitive Therapy and Research, 24, 233–249. doi: 10.1023/A:1005402310180 Katz, L. (2007). Domestic violence and vagal reactivity to peer provocation. Biological Psychology, 74, 154–164. doi: 10.1016/j.biopsycho.2005.10.010 Katz, L., & Gottman, J. M. (1995). Vagal tone protects children from marital conflict. Development and Psychopathology, 7, 83–92. Katz, L., & Gottman, J. M. (1996). Spillover effects of marital conflict: In search of parenting and coparenting mechanisms. In J. P. McHale & P. A. Cowan (Eds.), Understanding how family-level dynamics affect children's development: Studies of two-parent families (pp. 57– 76). San Francisco, CA: Jossey-Bass. Katz, L., & Gottman, J. M. (1997). Buffering children from marital conflict and dissolution. Journal of Clinical Child Psychology, 26, 157–171. doi: 10.1207/s15374424jccp2602_4 Katz, L., Hunter, E., & Klowden, A. (2008). Intimate partner violence and children's reaction to peer provocation: The moderating role of emotion coaching. Journal of Family Psychology, 22, 614–621. doi: 10.1037/a0012793 Katz, L., & Low, S. M. (2004). Marital violence, co-parenting and family-level processes, and children's adjustment. Journal of Family Psychology, 18, 372–382. Katz, L., & Woodin, E. M. (2002). Hostility, hostile detachment, and conflict engagement in marriages: Effects on child and family functioning. Child Development, 73, 636–651. doi: 10.1111/1467–8624.00428 Kelley, M. L., & Fals-Stewart, W. (2002). Couples-versus individual-based therapy for alcohol and drug abuse: Effects on children's psychosocial functioning. Journal of Consulting and Clinical Psychology, 70, 417–427. doi: 10.1037//0022–006X.70.2.417 Kelly, R. J., & El-Sheikh, M. (2013). Longitudinal relations between marital aggression and

children's sleep: The role of emotional insecurity. Journal of Family Psychology, 27, 282– 292. doi: 10.1037/a0031896 Kennedy, A. C., Bybee, D., Sullivan, C. M., & Greeson, M. (2010). The impact of family and community violence on children's depression trajectories: Examining the interactions of violence exposure, family social support, and gender. Journal of Family Psychology, 24, 197– 207. doi: 10.1037/a0018787 Kerig, P. K. (1995). Triangles in the family circle: Effects of family structure on marriage, parenting, and child adjustment. Journal of Family Psychology, 9, 28–43. doi: 10.1037/0893– 3200.9.1.28 Kerig, P. K. (1996). Assessing the links between interparental conflict and child adjustment: The conflicts and problem-solving scales. Journal of Family Psychology, 10, 454–473. doi: 10.1037/0893–3200.10.4.454 Kerig, P. K. (1998). Gender and appraisals as mediators of adjustment in children exposed to interparental violence. Journal of Family Violence, 13, 345–363. doi: 10.1023/A:1022871102437 Kerig, P. K. (2001). Children's coping with interparental conflict. In J. H. Grych & F. D. Fincham (Eds.), Interparental conflict and child development: Theory, research, and applications (pp. 213–245). New York, NY: Cambridge University Press. doi: 10.1017/CBO9780511527838.010 Kerig, P. K. (2005). Revisiting the construct of boundary dissolution: A multidimensional perspective. Journal of Emotional Abuse, 5, 5–42. doi: 10.1300/J135v05n02_02 Kerig, P. K., Fedorowicz, A. E., Brown, C. A., Patenaude, R. L., & Warren, M. (1998). When warriors are worriers: Gender and children's coping with interparental violence. Journal of Emotional Abuse, 1, 89–114. Kim, K. L., Jackson, Y., Conrad, S. M., & Hunter, H. L. (2008). Adolescent report of interparental conflict: The role of threat and self-blame appraisal on adaptive outcome. Journal of Child and Family Studies, 17, 735–751. Kinsfogel, K. M., & Grych, J. H. (2004). Interparental conflict and adolescent dating relationships: Integrating cognitive, emotional, and peer influences. Journal of Family Psychology, 18, 505–515. doi: 10.1037/0893–3200.18.3.505 Kitzmann, K. M., Gaylord, N. K., Holt, A. R., & Kenny, E. D. (2003). Child witnesses to domestic violence: A meta-analytic review. Journal of Consulting and Clinical Psychology, 71, 339–352. doi: 10.1037/0022–006X.71.2.339 Knous-Westfall, H. M., Ehrensaft, M. K., MacDonell, K. W., & Cohen, P. (2012). Parental intimate partner violence, parenting practices, and adolescent peer bullying: a prospective study. Journal of Family and Child Studies, 21, 754–766. doi: 10.1007/s10826-011-9528-2

Koss, K. J., George, M. R., Bergman, K. N., Cummings, E. M., Davies, P. T., & Cicchetti, D. (2011). Understanding children's emotional processes and behavioral strategies in the context of marital conflict. Journal of Experimental Child Psychology, 109, 336–352. Koss, K. J., George, M. R., Davies, P. T., Cicchetti, D., Cummings, E. M., & Sturge-Apple, M. L. (2013). Patterns of children's adrenocortical reactivity to interparental conflict and associations with child adjustment: A growth mixture modeling approach. Developmental Psychology, 49, 317–326. doi: 10.1037/a0028246 Kouros, C. D., Cummings, E. M., & Davies, P. T. (2010). Early trajectories of interparental conflict and externalizing problems as predictors of social competence in preadolescence. Development and Psychopathology, 22, 527–537. doi: 10.1017/S0954579410000258 Kouros, C. D., & El-Sheikh, M. (2015). Daily mood and sleep: Reciprocal relations and links with adjustment problems. Journal of Sleep Research, 24, 24–31. doi: 10.1111/jsr.12226 Kouros, C. D., Merrilees, C. E., & Cummings, E. M. (2008). Marital conflict and children's emotional security in the context of parental depression. Journal of Marriage and Family, 70, 684–697. doi: 10.1111/j.1741–3737.2008.00514.x Kouros, C. D., Papp, L. M., Goeke-Morey, M. C., & Cummings, E. M. (2014). Spillover between marital quality and parent-child relationship quality: Parental depressive symptoms as moderators. Journal of Family Psychology, 28, 315–325. Krishnakumar, A., & Buehler, C. (2000). Interparental conflict and parenting behaviors: A meta-analytic review. Family Relations: An Interdisciplinary Journal of Applied Family Studies, 49, 25–44. doi: 10.1111/j.1741–3729.2000.00025.x Lam, W. K., Fals-Stewart, W., & Kelley, M. L. (2008). Effects of parent skills training with behavioral couples therapy for alcoholism on children: A randomized clinical pilot trial. Addictive Behaviors, 33, 1076–80. doi: 10.1016/j.addbeh.2008.04.002 Lamb, M. E., Sternberg, K. J., & Thompson, R. A. (1997). The effects of divorce and custody arrangements on children's behavior, development, and adjustment. Family & Conciliation Courts Review, 35, 393–404. doi: 10.1111/j.174–1617.1997.tb00482.x Laumakis, M. A., Margolin, G., & John, R. S. (1998). The emotional, cognitive and coping responses of preadolescent children to different dimensions of marital conflict. In G. W. Holden, R. Geffner, & E. N. Jouriles (Eds.), Children exposed to marital violence: Theory, research, and applied issues (pp. 257–288). Washington, DC: American Psychological Association. doi: 10.1037/10257–008 Laursen, B., & Collins, W. A. (2009). Parent-child relationships during adolescence. In R. M. Lerner & L. Steinberg (Eds.), Handbook of adolescent psychology (3rd ed.), Vol. 2: Contextual influences on adolescent development (pp. 3–42). Hoboken, NJ: Wiley. doi: 10.1002/9780470479193.adlpsy002002

Leary, A., & Katz, L. (2004). Coparenting, family-level processes, and peer outcomes: The moderating role of vagal tone. Development and Psychopathology, 16, 593–608. doi: 10.1017/S0954579404004687 Ledermann, T., Bodenmann, G., & Cina, A. (2007). The efficacy of the Couples Coping Enhancement Training (CCET) in improving relationship quality. Journal of Social and Clinical Psychology, 26, 940–959. doi: 10.1521/jscp.2007.26.8.940 Lee, C. M., Beauregard, C., & Bax, K. A. (2005). Child-related disagreements, verbal aggression, and children's internalizing and externalizing behavior problems. Journal of Family Psychology, 19, 237–245. doi: 10.1037/0893–3200.19.2.237 Levendosky, A. A., & Graham-Bermann, S. A. (2000). Behavioral observations of parenting in battered women. Journal of Family Psychology, 14, 80–94. doi: 10.1037/0893–3200.14.1.80 Levendosky, A. A., & Graham-Bermann, S. A. (2001). Parenting in battered women: The effects of domestic violence on women and their children. Journal of Family Violence, 16, 171–192. doi: 10.1023/A:1011111003373 Levendosky, A. A., Huth-Bocks, A., & Semel, M. A. (2002). Adolescent peer relationships and mental health functioning in families with domestic violence. Journal of Clinical Child and Adolescent Psychology, 31, 206–218. doi: 10.1207/153744202753604485 Levendosky, A. A., Huth-Bocks, A. C., Shapiro, D. L., & Semel, M. A. (2003). The impact of domestic violence on the maternal-child relationship and preschool-age children's functioning. Journal of Family Psychology, 17, 275–287. doi: 10.1037/0893–3200.17.3.275 Levendosky, A. A., Leahy, K. L., Bogat, G., Davidson, W. S., & von Eye, A. (2006). Domestic violence, maternal parenting, maternal mental health, and infant externalizing behavior. Journal of Family Psychology, 20, 544–552. doi: 10.1037/0893–3200.20.4.544 Levendosky, A. A., Lynch, S. M., & Graham-Bermann, S. A. (2000). Mothers' perceptions of the impact of woman abuse on their parenting. Violence Against Women, 6, 247–271. doi: 10.1177/10778010022181831 Levy, G. D., & Fivush, R. (1993). Scripts and gender: A new approach for examining genderrole development. Developmental Review, 13, 126–146. doi: 10.1006/drev.1993.1006 Lieberman, A. F., Ghosh Ippen, C., & Van Horn, P. (2006). Child-parent psychotherapy: 6month follow-up of a randomized controlled trial. Journal of the American Academy of Child & Adolescent Psychiatry, 45, 913–918. Lieberman, A. F., Van Horn, P., & Ghosh Ippen, C. (2005). Toward evidence-based treatment: Child-parent psychotherapy with preschoolers exposed to marital violence. Journal of the American Academy of Child and Adolescent Psychiatry, 44, 1241–1248. Linder, J., & Collins, W. (2005). Parent and peer predictors of physical aggression and conflict

management in romantic relationships in early adulthood. Journal of Family Psychology, 19, 252–262. doi: 10.1037/0893–3200.19.2.252 Litrownik, A. J., Newton, R., Hunter, W. M., English, D., & Everson, M. D. (2003). Exposure to family violence in young at-risk children: A longitudinal look at the effects of victimization and witnessed physical and psychological aggression. Journal of Family Violence, 18, 59–73. doi: 10.1023/A:1021405515323 Lucas-Thompson, R. G. (2012). Associations of marital conflict with emotional and physiological stress: Evidence for different patterns of dysregulation. Journal of Research on Adolescence, 22, 704–721. doi: 10.1111/j.1532–7795.2012.00818.x Maccoby, E. E., & Mnookin, R. H. (1992). Dividing the child: Social and legal dilemmas of custody. Cambridge, MA: Harvard University Press. Mahoney, A., Boggio, R. M., & Jouriles, E. N. (1996). Effects of verbal marital conflict on subsequent mother–son interactions in a child clinical sample. Journal of Clinical Child Psychology, 25, 262–271. doi: 10.1207/s15374424jccp2503_2 Mann, B. J., & Gilliom, L. A. (2004). Emotional security and cognitive appraisals mediate the relationship between parents' marital conflict and adjustment in older adolescents. Journal of Genetic Psychology: Research and Theory on Human Development, 165, 250–271. doi: 10.3200/GNTP.165.3.250–271 Marcus, N., Lindahl, K. M., & Malik, N. M. (2001). Interparental conflict, children's social cognitions, and child aggression: A test of a mediational model. Journal of Family Psychology, 15, 315–333. doi: 10.1037/0893–3200.15.2.315 Margolin, G. (1981). Behavior exchange in happy and unhappy marriages: A family cycle perspective. Behavior Therapy, 12, 329–343. doi: 10.1016/S0005–7894(81)80122–0 Margolin, G., & Vickerman, K. A. (2007). Post-traumatic stress in children and adolescents exposed to family violence: I. Overview and issues. Professional Psychology, Research and Practice, 38, 620–628. doi: 10.1037/0735–7028.38.6.620 Margolin, G., Vickerman, K. A., Oliver, P. H., & Gordis, E. B. (2010). Violence exposure in multiple interpersonal domains: Cumulative and differential effects. Journal of Adolescent Health, 47, 198–205. doi: 10.1016/j.jadohealth.2010.01.020 Margolin, G., Vickerman, K. A., Ramos, M. C., Serrano, S., Gordis, E. B., Iturralde, E., … Spies, L. A. (2009). Youth exposed to violence: Stability, co-occurrence, and context. Clinical Child and Family Psychology Review, 12, 39–54. doi: 10.1007/s10567–009–0040–9 Martel, M. M., Nikolas, M., Jernigan, K., Friderici, K., & Nigg, J. T. (2012). Diversity in pathways to common childhood disruptive behavior disorders. Journal of Abnormal Child Psychology, 40, 1223–1236. doi: 10.1007/s10802–012–9646–3

Martinez-Torteya, C., Bogat, G., von Eye, A., & Levendosky, A. A. (2009). Resilience among children exposed to domestic violence: The role of risk and protective factors. Child Development, 80, 562–577. doi: 10.1111/j.1467–8624.2009.01279.x McCoy, K., Cummings, E., & Davies, P. T. (2009). Constructive and destructive marital conflict, emotional security and children's prosocial behavior. Journal of Child Psychology and Psychiatry, 50, 270–279. doi: 10.1111/j.1469–7610.2008.01945.x McDonald, R., & Grych, J. H. (2006). Young children's appraisals of interparental conflict: Measurement and links with adjustment problems. Journal of Family Psychology, 20, 88–99. doi: 10.1037/0893–3200.20.1.88 McDonald, R., Jouriles, E. N., & Minze, L. C. (2011). Interventions for young children exposed to intimate partner violence. In S. A. Graham-Bermann & A. A. Levendosky (Eds.), How intimate partner violence affects children: Developmental research, case studies, and evidence-based intervention (pp. 109–131). Washington, DC: American Psychological Association. doi: 10.1037/12322–006 McDonald, R., Jouriles, E. N., Ramisetty-Mikler, S., Caetano, R., & Green, C. E. (2006). Estimating the number of American children living in partner-violent families. Journal of Family Psychology, 20, 137–142. doi: 10.1037/0893–3200.20.1.137 McDonald, R., Jouriles, E. N., Rosenfield, D., & Leahy, M. M. (2012). Children's questions about interparent conflict and violence: What's a mother to say? Journal of Family Psychology, 26, 95–104. doi: 10.1037/a0026122 McDonald, R., Jouriles, E. N., & Skopp, N. A. (2006). Reducing conduct problems among children brought to women's shelters: Intervention effects 24 months following termination of services. Journal of Family Psychology, 20, 127–136. McDonald, R., Jouriles, E. N., Tart, C. D., & Minze, L. C. (2009). Children's adjustment problems in families characterized by men's severe violence toward women: Does other family violence matter? Child Abuse & Neglect, 33, 94–101. doi: 10.1016/j.chiabu.2008.03.005 McEachern, A. D., Fosco, G. M., Dishion, T. J., Shaw, D. S., Wilson, M. N., & Gardner, F. (2013). Collateral benefits of the Family Check-Up in early childhood: Primary caregivers' social support and relationship satisfaction. Journal of Family Psychology, (Advance online publication), 0–11. doi: 10.1037/a0031485 McHale, J. P., Johnson, D., and Sinclair, R. (1999). Family dynamics, preschoolers' family representations, and preschool peer relationships. Early Education and Development, 10, 373–401. McHale, J. P., Kuersten, R., & Lauretti, A. (1996). New directions in the study of family-level dynamics during infancy and early childhood. In J. P. McHale & P. A. Cowan (Eds.),

Understanding how family-level dynamics affect children's development: Studies of twoparent families (pp. 5–26). San Francisco, CA: Jossey-Bass. McHale, J. P., & Rasmussen, J. L. (1998). Coparental and family group-level dynamics during infancy: Early family precursors of child and family functioning during preschool. Development and Psychopathology, 10, 39–59. doi: 10.1017/S0954579498001527 McLoyd, V. C., & Smith, J. (2002). Physical discipline and behavior problems in African American, European American, and Hispanic children: Emotional support as a moderator. Journal of Marriage and Family, 64(1), 40–53. doi: 10.1111/j.1741–3737.2002.00040.x Metalsky, G. I., Joiner, T. E., Hardin, T. S., & Abramson, L. Y. (1993). Depressive reactions to failure in a naturalistic setting: A test of the hopelessness and self-esteem theories of depression. Journal of Abnormal Psychology, 102, 101–109. doi: 10.1037/0021– 843X.102.1.101 Mileva-Seitz, V. V., Kennedy, J. J., Atkinson, L. L., Steiner, M. M., Levitan, R. R., Matthews, S. G., …Fleming, A. S. (2011). Serotonin transporter allelic variation in mothers predicts maternal sensitivity, behavior and attitudes toward 6-month-old infants. Genes, Brain & Behavior, 10, 325–333. doi: 10.1111/j.1601–183X.2010.00671.x Minuchin, S. (1974). Families & family therapy. Oxford, UK: Harvard University Press. Moore, C. G., Probst, J. C., Tompkins, M., Cuffe, S., & Martin, A. B. (2007). The prevalence of violent disagreements in US families: Effects of residence, race/ethnicity, and parental stress. Pediatrics, 119 (Suppl. 1), S68–S76. doi: 10.1542/peds.2006–2089K Moore, G. A. (2010). Parent conflict predicts infants' vagal regulation in social interaction. Development and Psychopathology, 22, 23–33. doi: 10.1017/S095457940999023X Mrug, S., Loosier, P. S., & Windle, M. (2008). Violence exposure across multiple contexts: Individual and joint effects on adjustment. American Journal of Orthopsychiatry, 78, 70–84. doi: 10.1037/0002–9432.78.1.70 Mrug, S., & Windle, M. (2010). Prospective effects of violence exposure across multiple contexts on early adolescents internalizing and externalizing problems. Journal of Child Psychology and Psychiatry, 51, 953–961. doi: 10.1111/j.1469–7610.2010.02222.x Murphy, C. C., Schei, B., Myhr, T. L., & DuMont, J. (2001). Abuse: A risk factor for low birthweight? A systematic review and meta-analysis. Canadian Medical Association Journal, 164, 1567–1572. Nicolotti, L., El-Sheikh, M., & Whitson, S. M. (2003). Children's coping with marital conflict and their adjustment and physical health: Vulnerability and protective functions. Journal of Family Psychology, 17, 315–326. doi: 10.1037/0893–3200.17.3.315 Obradović, J., Bush, N. R., & Boyce, W. (2011). The interactive effect of marital conflict and

stress reactivity on externalizing and internalizing symptoms: The role of laboratory stressors. Development and Psychopathology, 23, 101–114. doi: 10.1017/S0954579410000672 O'Brien, M., Bahadur, M. A., Gee, C., Balto, K., & Erber, S. (1997). Child exposure to marital conflict and child coping responses as predictors of child adjustment. Cognitive Therapy and Research, 21, 39–59. doi: 10.1023/A:1021816225846 O'Brien, M., Margolin, G., & John, R. S. (1995). Relation among marital conflict, child coping, and child adjustment. Journal of Clinical Child Psychology, 24, 346–361. doi: 10.1207/s15374424jccp2403_12 O'Donnell, E. H., Moreau, M., Cardemil, E. V., & Pollastri, A. (2010). Interparental conflict, parenting, and childhood depression in a diverse urban population: The role of general cognitive style. Journal of Youth and Adolescence, 39, 12–22. doi: 10.1007/s10964–008– 9357–9 O'Leary, S. G., & Vidair, H. B. (2005). Marital adjustment, child-rearing disagreements, and overreactive parenting: Predicting child behavior problems. Journal of Family Psychology, 19, 208–216. doi: 10.1037/0893–3200.19.2.208 Oltmanns, T. F., Broderick, J. E., & O'Leary, K. (1977). Marital adjustment and the efficacy of behavior therapy with children. Journal of Consulting and Clinical Psychology, 45, 724–729. doi: 10.1037/0022–006X.45.5.724 Owen, M., & Cox, M. J. (1997). Marital conflict and the development of infant–parent attachment relationships. Journal of Family Psychology, 11, 152–164. doi: 10.1037/0893– 3200.11.2.152 Pajares, E. (2006). Self-efficacy during childhood and adolescence. In E. Pajares & T. Urban (Eds.), Self-efficacy beliefs of adolescents (pp. 307–337). Greenwich, CT: Information Age. Papp, L. M., Cummings, E., & Goeke-Morey, M. C. (2002). Marital conflicts in the home when children are present versus absent. Developmental Psychology, 38, 774–783. doi: 10.1037/0012–1649.38.5.774 Patterson, G. R. (1982). Coercive family process. Eugene, OR: Castalia. Patterson, G. R., Cobb, J. A., & Ray, R. S. (1973). A social engineering technology for retraining families of aggressive boys. In H. E. Adams & I. P. Unikel (Eds.), Issues and trends in behavior therapy. New York, NY: McGraw-Hill. Pauli-Pott, U., & Beckmann, D. (2007). On the association of interparental conflict with developing behavioral inhibition and behavior problems in early childhood. Journal of Family Psychology, 21, 529–532. doi: 10.1037/0893–3200.21.3.529 Peris, T. S., Goeke-Morey, M. C., Cummings, E., & Emery, R. E. (2008). Marital conflict and support seeking by parents in adolescence: Empirical support for the parentification construct.

Journal of Family Psychology, 22, 633–642. doi: 10.1037/a0012792 Peterson, J. L., & Zill, N. (1986). Marital disruption, parent-child relationships, and behavior problems in children. Journal of Marriage and the Family, 48, 295–307. doi: 10.2307/352397 Piotrowski, C. C., Tailor, K., & Cormier, D. C. (2014). Siblings exposed to intimate partner violence: Linking sibling relationship quality & child adjustment problems. Child Abuse & Neglect, 38, 123–134. doi: 10.1016/j.chiabu.2013.08.005 Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74, 116–143. doi: 10.1016/j.biopsycho.2006.06.009 Porter, B., & O'Leary, K. (1980). Marital discord and childhood behavior problems. Journal of Abnormal Child Psychology, 8, 287–295. doi: 10.1007/BF00916376 Pynoos, R. S., Steinberg, A. M., Ornitz, E. M., & Goenjian, A. K. (1997). Issues in the developmental neurobiology of traumatic stress. In R. Yehuda & A. C. McFarlane (Eds.), Psychobiology of posttraumatic stress disorder (pp. 176–193). New York, NY: New York Academy of Sciences. Reid, W. J., & Crisafulli, A. (1990). Marital discord and child behavior problems: A metaanalysis. Journal of Abnormal Child Psychology, 18, 105–117. doi: 10.1007/BF00919459 Repetti, R. L., Robles, T. F., & Reynolds, B. (2011). Allostatic processes in the family. Development and Psychopathology, 23, 921–938. doi: 10.1017/S095457941100040X Rhoades, K. A. (2008). Children's responses to interparental conflict: A meta-analysis of their associations with child adjustment. Child Development, 79, 1942–1956. doi: 10.1111/j.1467– 8624.2008.01235.x Riggs, D. S., & O'Leary, K. (1996). Aggression between heterosexual dating partners: An examination of a causal model of courtship aggression. Journal of Interpersonal Violence, 11, 519–540. doi: 10.1177/088626096011004005 Rogers, M., & Holmbeck, G. N. (1997). Effects of interparental aggression on children's adjustment: The moderating role of cognitive appraisal and coping. Journal of Family Psychology, 11, 125–130. doi: 10.1037/0893–3200.11.1.125 Rutter, M. (1971). Parent–child separation: Psychological effects on the children. Journal of Child Psychology and Psychiatry, 12, 233–260. Rutter, M. (1979). Protective factors in children's responses to stress and disadvantage. Annals of the Academy of Medicine Singapore, 8, 324. Rutter, M. (1987). Psychosocial resilience and protective mechanisms. American Journal of Orthopsychiatry, 57, 316–331. doi: 10.1111/j.1939–0025.1987.tb03541.x

Salomon, K., Matthews, K. A., & Allen, M. T. (2000). Patterns of sympathetic and parasympathetic reactivity in a sample of children and adolescents. Psychophysiology, 37, 842–849. doi: 10.1017/S0048577200990607 Saltzman, K. M., Holden, G. W., & Holahan, C. J. (2005). The psychobiology of children exposed to marital violence. Journal of Clinical Child and Adolescent Psychology, 34, 129– 139. doi: 10.1207/s15374424jccp3401_12 Satir, V. M. (1964). Conjoint family therapy. Palo Alto, CA: Science and Behavior Books. Schermerhorn, A. C., Cummings, E., DeCarlo, C. A., & Davies, P. T. (2007). Children's influence in the marital relationship. Journal of Family Psychology, 21, 259–269. doi: 10.1037/0893–3200.21.2.259 Schwarz, B., Stutz, M., & Ledermann, T. (2012). Perceived interparental conflict and early adolescents' friendships: The role of attachment security and emotion regulation. Journal of Youth and Adolescence, 41, 1240–1252. doi: 10.1007/s10964-9769-4 Shamir, H., Cummings, E., Davies, P. T., & Goeke-Morey, M. C. (2005). Children's reactions to marital conflict in Israel and in the United States. Parenting: Science and Practice, 5, 371– 386. doi: 10.1207/s15327922par0504_4 Shelton, K. H., & Harold, G. T. (2007). Marital conflict and children's adjustment: The mediating and moderating role of children's coping strategies. Social Development, 16, 497– 512. doi: 10.1111/j.1467–9507.2007.00400.x Shelton, K. H., & Harold, G. T. (2008). Pathways between interparental conflict and adolescent psychological adjustment: Bridging links through children's cognitive appraisals and coping strategies. Journal of Early Adolescence, 28, 555–582. doi: 10.1177/0272431608317610 Shifflett-Simpson, K., & Cummings, E. (1996). Mixed message resolution and children's responses to interadult conflict. Child Development, 67, 437–448. doi: 10.2307/1131825 Skopp, N. A., McDonald, R., Jouriles, E. N., & Rosenfield, D. (2007). Partner aggression and children's externalizing problems: Maternal and partner warmth as protective factors. Journal of Family Psychology, 21, 459–467. doi: 10.1037/0893–3200.21.3.459 Skopp, N. A., McDonald, R., Manke, B., & Jouriles, E. N. (2005). Siblings in domestically violent families: Experiences of interparent conflict and adjustment problems. Journal of Family Psychology, 19, 324–333. doi: 10.1037/0893–3200.19.2.324 Slep, A., & O'Leary, S. G. (2001). Examining partner and child abuse: Are we ready for a more integrated approach to family violence? Clinical Child and Family Psychology Review, 4, 87–107. doi: 10.1023/A:1011319213874 Slep, A., & O'Leary, S. G. (2009). Distinguishing risk profiles among parent-only, partner-

only, and dually perpetrating physical aggressors. Journal of Family Psychology, 23, 705– 716. doi: 10.1037/a0016474 Smith, C. A., Ireland, T. O., Park, A., Elwyn, L., & Thornberry, T. P. (2011). Intergenerational continuities and discontinuities in intimate partner violence: A two-generational prospective study. Journal of Interpersonal Violence, 26, 3720–3752. doi: 10.1177/0886260511403751 Snyder, D. K., Klein, M. A., Gdowski, C. L., Faulstich, C., & LaCombe, J. (1988). Generalized dysfunction in clinic and nonclinic families: A comparative analysis. Journal of Abnormal Child Psychology, 16, 97–109. Spaccarelli, S., Coatsworth, J., & Bowden, B. (1995). Exposure to serious family violence among incarcerated boys: Its association with violent offending and potential mediating variables. Violence and Victims, 10, 163–182. Steinberg, L., & Avenevoli, S. (2000). The role of context in the development of psychopathology: A conceptual framework and some speculative propositions. Child Development, 71, 66–74. doi: 10.1111/1467–8624.00119 Stocker, C., Ahmed, K., & Stall, M. (1997). Marital satisfaction and maternal emotional expressiveness: Links with children's sibling relationships. Social Development, 6, 373–385. doi: 10.1111/j.1467–9507.1997.tb00112.x Stocker, C. M., & Youngblade, L. (1999). Marital conflict and parental hostility: Links with children's sibling and peer relationships. Journal of Family Psychology, 13, 598–609. doi: 10.1037/0893–3200.13.4.598 Stoneman, Z., & Brody, G. H. (1993). Sibling temperaments, conflict, warmth, and role asymmetry. Child Development, 64, 1786–1800. doi: 10.2307/1131469 Stroud, C. B., Meyers, K. M., Wilson, S., & Durbin, C. E. (2015). Marital quality spillover and young children's adjustment: Evidence for dyadic and triadic parenting as mechanisms. Journal of Clinical Child & Adolescent Psychology, 44, 800–813, doi: 10.1080/15374416.2014.900720 Sturge-Apple, M. L., Cicchetti, D., Davies, P. T., & Suor, J. H. (2012). Differential susceptibility in spillover between interparental conflict and maternal parenting practices: Evidence for OXTR and 5-HTT genes. Journal of Family Psychology, 26, 431–442. doi: 10.1037/a0028302 Sturge-Apple, M. L., Skibo, M. A., & Davies, P. T. (2012). Impact of parental conflict and emotional abuse on children and families. Partner Abuse, 3, 379–400. Sullivan, C. M., & Bybee, D. I. (1999). Reducing violence using community-based advocacy for women with abusive partners. Journal of Consulting and Clinical Psychology, 67, 43–53. Sullivan, C. M., Bybee, D. I., & Allen, N. E. (2002). Findings from a community-based

program for battered women and their children. Journal of Interpersonal Violence, 17, 915– 936. doi: 10.1177/0886260502017009001 Tailor, K., Stewart-Tufescu, A., & Piotrowski, C. (2015). Children exposed to intimate partner violence: Influences of parenting, family distress, and siblings. Journal of Family Psychology, 29, 29–38. doi: 10.1037/a0038584 Towle, C. C. (1931). The evaluation and management of marital situation in foster homes. American Journal of Orthopsychiatry, 1, 271–283. doi: 10.1111/j.1939–0025.1931.tb04821.x Trickey, D., Siddaway, A. P., Meiser-Stedman, R., Serpell, L., & Field, A. P. (2012). A metaanalysis of risk factors for post-traumatic stress disorder in children and adolescents. Clinical psychology review, 32, 122–38. doi: 10.1016/j.cpr.2011.12.001 Tschann, J. M., Flores, E., Pasch, L. A., & Marin, B. V. (1999). Assessing interparental conflict: Reports of parents and adolescents in European American and Mexican American families. Journal of Marriage and the Family, 61(2), 269–283. Ulbricht, J. A., Ganiban, J. M., Button, T. M., Feinberg, M., Reiss, D., & Neiderhiser, J. M. (2013). Marital adjustment as a moderator for genetic and environmental influences on parenting. Journal of Family Psychology, 27, 42–52. doi: 10.1037/a0031481 Wasserstein, S. B., & La Greca, A. M. (1996). Can peer support buffer against behavioral consequences of parental discord? Journal of Clinical Child Psychology, 25, 177–182. doi: 10.1207/s15374424jccp2502_6 Whisman, M. A., Beach, S. H., & Snyder, D. K. (2008). Is marital discord taxonic and can taxonic status be assessed reliably? Results from a national, representative sample of married couples. Journal of Consulting and Clinical Psychology, 76, 745–755. doi: 10.1037/0022– 006X.76.5.745 Wolfe, D. A., Crooks, C. V., Lee, V., McIntyre-Smith, A., & Jaffe, P. G. (2003). The effects of children's exposure to domestic violence: A meta-analysis and critique. Clinical Child and Family Psychology Review, 6, 171–187. doi: 10.1023/A:1024910416164 Wolfe, D. A., Wekerle, C., Scott, K., Straatman, A., & Grasley, C. (2004). Predicting abuse in adolescent dating relationships over 1 year: The role of child maltreatment and trauma. Journal of Abnormal Psychology, 113, 406–415. doi: 10.1037/0021–843X.113.3.406 Wood, B. L., Klebba, K. B., & Miller, B. D. (2000). Evolving the biobehavioral family model: The fit of attachment. Family Process, 39, 319–344. doi: 10.1111/j.1545–5300.2000.39305.x Wymbs, B. T., Pelham, W. R., Molina, B. G., Gnagy, E. M., Wilson, T. K., & Greenhouse, J. B. (2008). Rate and predictors of divorce among parents of youths with ADHD. Journal of Consulting and Clinical Psychology, 76, 735–744. doi: 10.1037/a0012719 Yates, T. M., Dodds, M. F., Sroufe, L., & Egeland, B. (2003). Exposure to partner violence and

child behavior problems: A prospective study controlling for child physical abuse and neglect, child cognitive ability, socioeconomic status, and life stress. Development and Psychopathology, 15, 199–218. doi: 10.1017/S0954579403000117 Yoo, J., & Huang, C. (2012). The effects of domestic violence on children's behavior problems: Assessing the moderating roles of poverty and marital status. Children and Youth Services Review, 34, 2464–2473. doi: 10.1016/j.childyouth.2012.09.014 Zeitlin, D., Dhanjal, T. T., & Colmsee, M. M. (1999). Maternal-foetal bonding: The impact of domestic violence on the bonding process between a mother and child. Archives of Women's Mental Health, 2, 183–189. doi: 10.1007/s007370050047

Chapter 13 Relational Aggression: A Developmental Psychopathology Perspective Dianna Murray-Close, David A. Nelson, Jamie M. Ostrov, Juan F. Casas, and Nicki R. Crick This chapter is dedicated to the memory and legacy of Dr. Nicki R. Crick, a pioneer in the study of relational aggression. The rich body of theoretical and empirical work covered in this chapter is a testament to the far-reaching influence of Crick's contributions to the field of developmental psychopathology. We thank Alexandra Sullivan, Chelsea Krisanda, Casey Gates, and Lacey Evans for their assistance in compiling references for this chapter. Preparation of this manuscript was facilitated by a grant from the National Science Foundation (0819148) to the first and last authors. Please direct all correspondence to the first author at 2 Colchester Ave., Department of Psychological Science, University of Vermont, Burlington, VT, 05405; [email protected]. INTRODUCTION DEFINING RELATIONAL AGGRESSION Historical Perspectives Forms and Functions of Aggression DEVELOPMENTAL CHANGE IN RELATIONAL AGGRESSION Mean Changes in Relational Aggression Developmental Manifestations of Relational Aggression Developmental Challenges in Assessment AGGRESSION AND GENDER Theoretical Perspectives Regarding Aggressive Girls Gender Differences in Relational Aggression RISK FACTORS: BIOBEHAVIORAL PROCESSES Genetic Risk Psychophysiological Stress Systems Additional Biological Factors RISK FACTORS: COGNITIVE AND EMOTIONAL PROCESSES The Crick and Dodge (1994) Reformulated Social Information Processing Model of Children's Adjustment

Toward an Integrated Gender-Linked Model of Aggression Emotion and Emotion Regulation RISK FACTORS: SOCIAL PROCESSES Parenting and Attachment Parental Beliefs and Expectations Interparental Conflict Sibling Relationships Peer Relationships Media Exposure DEVELOPMENTAL OUTCOMES: MALADAPTIVE CORRELATES Internalizing Problems Externalizing Problems Peer Problems Substance Use and Abuse Personality Pathology DEVELOPMENTAL OUTCOMES: POTENTIAL POSITIVE CORRELATES CULTURAL PERSPECTIVES RELATIONAL AGGRESSION INTERVENTIONS DEVELOPMENTAL PSYCHOPATHOLOGY PERSPECTIVES Developmental Tasks Multilevel Perspectives Equifinality and Multifinality FUTURE DIRECTIONS Understanding How Relational Aggression Relates to Antisocial Behaviors Moving Beyond “Mean Girls” Developing New Methods and Research Designs to Study Relational Aggression Culture and Context CONCLUSION REFERENCES

Introduction The development of aggressive behavior patterns has long captured the theoretical and empirical attention of developmental psychopathologists. Although much of this work has focused on physically (e.g., hitting, kicking) or overtly (i.e., physical aggression combined with verbal aggression, such as verbal insults) aggressive behavior, in recent years there has been significant interest in the development of relational aggression. Relational aggression, defined as behaviors intended to hurt or harm others using relationships as the vehicle of harm, includes behaviors such as spreading malicious gossip, socially excluding peers, and using the silent treatment (Crick & Grotpeter, 1995). The goal of this chapter is to evaluate the state of knowledge regarding relational aggression two decades after Crick and Grotpeter's (1995) seminal publication on the topic. First, we focus on conceptual challenges in defining relational aggression, normative and atypical developmental trajectories, and theoretical and empirical work regarding gender differences in such conduct. Second, we address potential mechanisms involved in the development of relational aggression, including biobehavioral, cognitive-emotional, and social processes. Third, we consider both positive and negative developmental outcomes associated with involvement in relational aggression. Fourth, we discuss implications of research on relational aggression for both cross-cultural research and interventions. Finally, we consider the significance of a developmental psychopathology perspective for the study of relational aggression as well as the most critical directions for future research in this area.

Defining Relational Aggression Historical Perspectives Numerous theoretical models and traditions have guided the study of the development of aggressive behavior (for reviews, see Crick & Zahn-Waxler, 2003; Ostrov & Godleski, 2010). A review of the historical record suggests that eminent scholars such as Gordon W. Allport were examining indirect aggression as early as 1941 (Allport, Bruner, & Jandorf, 1941). In 1961, Arnold H. Buss, in his classic “Psychology of Aggression,” noted that: From the aggressor's vantage point, the best model of aggression is one that avoids counterattack. Indirect aggression solves the problem by rendering it difficult to identify the aggressor. Indirect aggression may be verbal (spreading nasty gossip) or physical (a man sets fire to his neighbor's home). These examples illustrate the two ways in which aggression can be indirect. Gossip is indirect in that the victim is not present.…Damaging a person's possessions is indirect in that the victim is not hurt or injured, but objects associated with and valued by him are destroyed. (p. 8). Moreover, Buss (1961) cites Eron, Laulicht, and Walder (1956) and their “Rip Van Winkle research team” for the development of a peer rating instrument that includes an assessment of indirect aggression: “Who makes up stories and lies to get other children in trouble?” (p. 269). Much of the early work regarding nonphysical forms of aggression, including indirect

aggression, focused on gender differences in such conduct. Norma D. Feshbach (1969) conducted an empirical study of direct and indirect aggression and found that girls displayed significantly higher levels of indirect aggression than did boys. Nearly two decades later, Lagerspetz, Björkqvist, and Peltonen (1988) examined gender differences in aggressive behaviors among 11- and 12-year-old boys and girls in Finland. Using peer reports and individual interviews, these authors identified a 3-factor solution to children's responses to the question “what do they/I do when angry with another boy/girl in the class?” The first factor indicated “what we have called ‘indirect means,’ and includes circumventory behaviour that exploits social relations among peers to harm the person at whom the anger is directed” (Lagerspetz et al., 1988, p. 409). This factor included 10 items, such as “tells untruth behind back,” “says to others: ‘Let's not be with him/her,’” “starts being somebody else's friend in revenge,” “argues,” and “sulks” (p. 409). The second factor focused on “direct means of aggression,” and the third factor (two items) included “peaceful means of solving conflicts” (p. 409). It is clear that this conceptualization of indirect aggression was influenced by the writing of Arnold Buss (1961), which was appropriately cited by the authors (p. 404). In 1992, Björkqvist, Lagerspetz, and Kaukiainen further clarified: “Indirect aggression is a type of behaviour in which the perpetrator attempts to inflict pain in such a manner that he or she makes it seem as though there has been no intention to hurt at all. Accordingly, he or she is more likely to avoid counter aggression and, if possible, to remain unidentified” (p. 118). This definition mirrors recent descriptions of passive aggression, which is conceptually distinct from relational forms of aggression (e.g., Nelson, Springer, Nelson, & Bean, 2008). Moreover, unlike relational aggression, indirect aggression is unlikely to occur in young children (Björkqvist et al., 1992). In 1989, Robert Cairns and colleagues introduced the related term social aggression, which they defined as the “manipulation of group acceptance through alienation, ostracism, or character defamation” (Cairns, Cairns, Neckerman, Ferguson, & Gariépy, 1989, p. 323). As originally conceptualized by Cairns and colleagues, social aggression involves the peer group and thus may be different from indirect aggression, which may involve either dyadic or grouplevel transgressions. Moreover, unlike relational aggression, social aggression likely would not be displayed at younger ages when the importance of the larger peer group and social status is not yet established. In their 6-year longitudinal study examining the growth in aggressive behavior from fourth grade to early adolescence, Cairns and colleagues found that social aggression was highly representative of girls' aggressive behavior during early adolescence. Building on the concepts of indirect and social aggression, Crick and Grotpeter (1995) identified a relational form of aggression that they argued was relatively common among girls. They cited prior work on indirect and social aggression but suggested that the central element of girls' aggressive behavior patterns was the use of interpersonal relationships to harm others. More specifically, these authors argued that “girls' attempts to harm others would focus on relational issues and would include behaviors intended to significantly damage another child's friendships or feelings of inclusion by the peer group…harming others through purposeful manipulation and damage of their peer relationships” (p. 711). Rather than focusing on efforts to avoid detection or retribution, as was often highlighted in theoretical formulations of

indirect aggression, Crick and Grotpeter proposed a central role of children's social goals in their selection of aggressive behaviors. From this perspective, girls were expected to engage in relatively high levels of relational aggression because these behaviors reflect the relational issues that are characteristic of female peer groups. This initial study demonstrated that girls were not only more relationally aggressive than school-age boys but that, relative to their nonrelationally aggressive peers, children that engaged in high levels of relational aggression were at increased risk for serious adjustment problems (i.e., greater probability of peer rejection, loneliness, and symptoms of depression). In 1997, Galen and Underwood expanded on the definition of social aggression. They argued that “the construct of relational aggression may not capture all of the forms of aggression evident in girls' peer interactions. Negative facial expressions and gestures and subtle jabs at another's self-esteem may also be important features of girls' aggressive behavior.…Social ostracism or relationship manipulation may begin with rolling of eyes, tossing of hair, and turning away from a peer” (p. 590). Accordingly, nonverbal acts beyond the silent treatment (which was included in Crick and Grotpeter's original measure in 1995) were considered to be important behaviors for empirical investigation. In recent years, research on indirect, relational, and social aggression has flourished, and studies have documented that these constructs provide significant information regarding key developmental processes. However, considerable disagreement persists regarding the degree of overlap among these constructs (see Archer & Coyne, 2005). There are clearly some components of indirect aggression that are not shared by relational aggression. For instance, indirect physical aggression, wherein the identity of the perpetrator is unknown (e.g., placing gum on a peer's chair), has been considered to be an exemplar of indirect aggression (see Goldstein, Tisak, & Boxer, 2002). As these behaviors do not use relationships as the vehicle of harm, however, they would not be considered cases of relational aggression. Additionally, there are some components of social aggression that are not shared by relational aggression (e.g., nonverbal acts such as eye rolling). Methodologically, then, research instruments developed to assess each form of aggression are likely to capture somewhat different behavioral patterns. Perhaps more important, there are several reasons to expect distinct risk factors, developmental trajectories, and developmental outcomes associated with each of these defined subtypes of aggression. For instance, investment in relational goals, such as the establishment of close interpersonal relationships, may be most strongly associated with the enactment of relational forms of aggression. Developmentally, relational aggression may be observed at earlier ages than either indirect or social aggression (see the section titled “Developmental Manifestations of Relational Aggression”). Moreover, given the central role of damage to interpersonal relationships in relational forms of aggression, interdependent cultural contexts may be particularly relevant to the developmental outcomes associated with relational forms of aggression (see Kawabata, Crick, & Hamaguchi, 2010). Thus, despite the conceptual overlap among these subtypes of aggression, we believe that it is critically important for researchers to be precise in their use of terminology, to adopt terms and associated methods that are theoretically based, and to maintain the fidelity of the methodology of the original instrument

developers.

Forms and Functions of Aggression Although all aggressive behaviors involve actions intended to hurt, harm, or injure another person, developmental scholars have made distinctions between forms (i.e., the whats) and functions (i.e., the whys) of aggressive behaviors (e.g., Little, Jones, Henrich, & Hawley, 2003; Prinstein & Cillessen, 2003). Researchers have identified a number of distinct forms of aggression, including physical, relational, verbal, nonverbal, and electronic aggression. However, in the last two decades, the most common delineation in developmental psychopathology research has been between physical and nonphysical forms of aggression, with relational aggression being a key construct in most of these studies (e.g., Burnette & Reppucci, 2009; Crick, 1995; Kliewer, Dibble, Goodman, & Sullivan, 2012; Preddy & Fite, 2012). Physical aggression is defined as acts intended to hurt others via physical force or the threat of physical force (Crick & Grotpeter, 1995). Classic examples of physical aggression include behaviors such as hitting, pinching, kicking, biting, and taking objects (Crick & Grotpeter, 1995). In contrast, relational aggression is defined as behaviors that damage or threaten to damage close relationships to hurt, harm, or injure another person (Crick & Grotpeter, 1995). Relational aggression may be manifested in different ways during different developmental periods (see “Developmental Manifestations of Relational Aggression” section), but it includes both direct and indirect behaviors and may be displayed verbally or nonverbally (Nelson, Springer, et al., 2008). Prototypical examples of direct relational aggression include telling someone that they are not invited to a party (exclusion) or placing one's hand on a chair so another child cannot sit down. Indirect relational aggression includes gossip and rumor spreading. Physical and relational aggression have emerged as unique factors in several studies (e.g., Crick & Grotpeter, 1995), are theorized to be associated with different social goals (Crick & Grotpeter, 1995), and have been found to be associated with different risk factors and developmental outcomes (see Crick, Ostrov, & Kawabata, 2007; Ostrov & Godleski, 2010). Scholars have also distinguished between the functions of aggression. Although various terms are used within the psychological sciences, the two functions of aggression are typically referred to as proactive aggression (i.e., also known as instrumental, premeditated, or coldblooded) and reactive aggression (i.e., also known as hostile, impulsive, or hot-blooded; Vitaro, Gendreau, Tremblay, & Oligny, 1998). Proactive aggression is defined as goal-directed aggressive behavior; in contrast, reactive aggression is concurrently displayed in response to a real or perceived threat and is often motivated by hostility or anger (Vitaro et al., 1998). Despite some overlap in the display of these functions, evidence indicates that these subtypes of aggression are distinct factors (e.g., Poulin & Boivin, 2000). Moreover, proactive and reactive aggression have different theoretical underpinnings (i.e., social learning theory for proactive aggression and the frustration-aggression hypothesis for reactive aggression; Bandura, 1973; Berkowitz, 1962). Accordingly, unique developmental precursors and socialpsychological adjustment outcomes are related to proactive and reactive functions of

aggression (for a meta-analysis, see Card & Little, 2006). Despite the tradition in psychological research of examining forms and functions of aggression separately, more recent research has taken into account both forms and functions of aggression. There are two main methodological approaches to examining forms and functions of aggression. The first uses a two-dimensional combination approach in which the two forms (i.e., physical and relational) and the two functions (i.e., proactive and reactive) are crossed, yielding four subtypes of aggression: proactive physical, reactive physical, proactive relational, and reactive relational (e.g., Ostrov & Crick, 2007; Prinstein & Cillessen, 2003). This approach is purported to be ecologically valid (Underwood, 2003), and research adopting this method has demonstrated unique correlates of subtypes of aggression in early childhood (e.g., Ostrov, Murray-Close, Godleski, & Hart, 2013), middle childhood (e.g., Mathieson & Crick, 2010), adolescence (e.g., Prinstein & Cillessen, 2003), and emerging adulthood (e.g., Bailey & Ostrov, 2008). However, the two-dimensional combination approach is plagued by high levels of statistical overlap between proactive and reactive physical aggression as well as proactive and reactive relational aggression. For this reason, Little and colleagues introduced a structural equation modeling method to disentangle form and function by creating statistically pure form and function latent variables (Little et al., 2003). Several studies have now successfully used this approach with child and adolescent samples (e.g., Fite, Stauffacher, Ostrov, & Colder, 2008; Fite, Stoppelbein, Greening, & Gaertner, 2009; Murray-Close & Ostrov, 2009; see a related approach in Sijtsema, Ojanen, et al., 2010), and these models have revealed unique predictors and outcomes associated with the pure constructs. For example, in keeping with theory, hostility was associated with pure reactive, physical, and relational aggression but not with proactive aggression (Little et al., 2003). Other researchers, however, have suggested that functions reflect differences in severity, rather than type, of aggression (Marsee et al., 2014; Stickle, Marini, & Thomas, 2012). For instance, using person-centered analyses, Marsee and colleagues (2014) identified three groups of relationally aggressive adolescent girls in two samples (voluntary residential and involuntarily detained): nonaggressive, reactive relationally aggressive, and combined proactive/reactive relationally aggressive. Consistent with a severity (rather than typology) model, some evidence indicated that the combined group exhibited the most severe adjustment difficulties. Interestingly, these three groups emerged for girls but not for boys. The findings suggest that proactive relational aggression is unlikely to occur in the absence of reactive relational aggression and that girls who exhibit proactive relational aggression may be exhibiting aggressive behaviors that are more severe in nature. Future research is needed to examine the unique and shared developmental pathways that these forms and functions denote for individuals in various contexts.

Developmental Change in Relational Aggression Mean Changes in Relational Aggression

Evidence indicates that relational aggression emerges quite early in development; in fact, observational studies have documented these behaviors in children as young as 30 months old (Crick, Ostrov, Burr, et al., 2006). Moreover, relational aggression has been reported among the elderly in assisted living facilities (Trompetter, Scholte, & Westerhof, 2011), suggesting that these behaviors may occur with some frequency across the life span. However, a significant question for developmental psychopathologists is whether the frequency of relational aggression changes across development. As Underwood, Galen, and Paquette (2001) discussed, there has been considerable debate in the field regarding whether social and relational aggression are common, normative social behaviors or whether such conduct reflects a relatively atypical and maladaptive behavior pattern. As these researchers pointed out, the answer to this question may depend on the developmental period under consideration, highlighting the need for work identifying typical developmental trajectories of relational aggression. A number of theorists have argued that there will be important developmental differences in the use of relational aggression as a result of cognitive, social, and biological changes. In early theoretical discussions of this issue, Björkqvist et al. (1992) suggested that aggression is a relatively normative behavior, but as young children are limited by their social and cognitive skills, they must rely on unsophisticated forms of aggression, such as physical aggression (see also Lagerspetz & Björkqvist, 1994). As children get older and their social and cognitive abilities improve, however, they are hypothesized to replace these physically aggressive behaviors with more covert forms of aggression, such as indirect aggression (Lagerspetz & Björkqvist, 1994). The successful use of relationally aggressive behaviors, such as spreading a rumor about a peer, may require relatively advanced social-cognitive skills (e.g., theory of mind abilities) and verbal abilities (e.g., Bonica, Arnold, Fisher, Zeljo, & Yershova, 2003). In fact, findings from several past studies indicate that indirect and relational forms of aggression are related to heightened social-cognitive abilities broadly defined (Andreou, 2006; Kaukiainen et al., 1999; for review, see Sutton, Smith, & Swettenham, 1999). Relational aggression in preschoolers has been found to be related to expressive and receptive language abilities (Bonica et al., 2003; Hawley, 2003b; although see Côté, Vaillancourt, Barker, Nagin, & Tremblay, 2007; Estrem, 2005; Kikas, Peets, Tropp, & Hinn, 2009), deception skills (Ostrov, 2006; Ostrov, Ries, Stauffacher, Godleski, & Mullins, 2008), theory of mind (Renouf et al., 2010), and, in females, moral maturity (Hawley, 2003b). In addition, although somewhat sophisticated relationally aggressive acts (e.g., secret spreading) have been documented in children as young as 3 years of age, the identity of the perpetrator is known and the behaviors are based on the here and now rather than events in the past (Ostrov, Woods, Jansen, Casas, & Crick, 2004). Taken together, these findings suggest that advances in social-cognitive skills across development may at times promote relationally aggressive behavior; however, it is important to note that improved skills in these areas may be most likely to predict aggressive conduct in subsets of youth who exhibit impairments in other cognitive abilities (see Crick & Dodge, 1999). Significant social changes during childhood and adolescence may also lead to increased use of relational forms of aggression. Because relational aggression targets relationships as the

vehicle of harm, these behaviors may be most effective in the context of established social networks (Archer & Coyne, 2005) and close, intimate friendships (e.g., Murray-Close, Ostrov, & Crick, 2007), both of which become more common during adolescence (e.g., Berndt, 1996). In recent years, theorists have argued that relational aggression may be an effective strategy to establish and maintain high status in the peer group, particularly during adolescence (Cillessen & Mayeux, 2004; see the “Developmental Outcomes: Potential Positive Correlates” section). As youth approach adolescence, they increasingly prioritize high peer status over other domains (LaFontana & Cillessen, 2010); thus, relational aggression may be perceived as a particularly important tool during this developmental period. Additionally, as children's concerns about the social repercussions of behaviors increase, they may be inclined to select aggressive strategies that are less likely to lead to negative social sanctions, such as social or relational aggression (e.g., Karriker-Jaffe, Foshee, Ennett, & Suchindran, 2008). Finally, some researchers have argued that biological changes related to puberty may predict increases in relationally aggressive conduct (e.g., Kistner et al., 2010). Puberty is accompanied by a host of physical changes, including increases in reproductive hormones, such as testosterone. Reproductive hormones, in turn, may facilitate aggressive behaviors in the context of competition for mates (Archer, 2006). Although little work has examined the role of reproductive hormones in relationally aggressive behaviors, Sánchez-Martín et al. (2011) provided initial evidence that testosterone was positively associated with relational aggression in 9-year-old males and females. Pubertal changes may also facilitate relational aggression because they lead to competition for access to sexual partners; in fact, several researchers have argued that relational aggression may be used as a strategy to derogate sexual rivals (Kistner et al., 2010; Pellegrini & Long, 2003; White, Gallup, & Gallup, 2010). Consistent with this perspective, dating popularity is associated with relational aggression, particularly among girls (Pellegrini & Long, 2003), and in one study, earlier age of first intercourse was related to heightened self-reported indirect aggression (White et al., 2010). Finally, early puberty may function as a transitional stressor that promotes aggressive conduct; for instance, Susman et al. (2007) found that early puberty was associated with higher relational aggression in girls. Similarly, Hemphill et al. (2010) found that pubertal stage was positively associated with social aggression among younger, but not older, adolescents. Consistent with the perspective that cognitive, social, and biological changes across early and middle childhood promote relational forms of aggression, an emerging body of work indicates that relational aggression becomes increasingly common as youth approach adolescence (e.g., Karriker-Jaffe et al., 2013), particularly among girls (Kawabata, Tseng, Murray-Close, & Crick, 2012; Kistner et al., 2010; Murray-Close et al., 2007). For instance, Murray-Close et al. (2007) reported that, among girls only, relational aggression increased in frequency over the course of one calendar year during late elementary school. In their person-centered analyses of trajectories of indirect aggression from ages 2 to 8, Côté et al. (2007) found that most youth exhibited low, stable levels of indirect aggression. However, a subset of youth evidenced increases in their indirect aggression across the course of the study, and girls outnumbered boys in this trajectory group. Interestingly, some of these children also exhibited a pattern of decreasing physical aggression, providing partial support for the hypothesis that youth will

replace their physically aggressive behaviors with relational or indirect aggression as their cognitive and social skills mature (see also J. L. Miller, Vaillancourt, & Boyle, 2009; although see Vaillancourt, Brendgen, Boivin, & Tremblay, 2003). However, not all researchers have documented increases in relational aggression during childhood and early adolescence. For instance, Spieker et al. (2012) reported that girls exhibited fairly stable levels of relational aggression from third to sixth grade, whereas boys exhibited declines in their involvement in such conduct. Additionally, Underwood et al. (2009) found that social aggression decreased in frequency from third to seventh grade, and personcentered analyses indicated that the sample was characterized by one group of children who displayed stable and low levels of social aggression and another group that initially engaged in relatively high levels of social aggression but exhibited decreases in such conduct across the study. One potential explanation regarding these mixed findings is that relational aggression becomes increasingly common as youth approach preadolescence but then declines in frequency following the transition to adolescence (Karriker-Jaffe et al., 2008). In fact, at least some of the cognitive, social, and biological processes hypothesized to underlie increases in the use of relational aggression are likely to have stabilized or waned by late adolescence. For instance, the social changes thought to promote increased involvement in relational aggression (e.g., heightened focus on peers and social status) may become less common as friendships become more autonomous in nature (Karriker-Jaffe et al., 2008). Similarly, as youth get older, biological transitions, such as puberty, may be less stressful and thus may be less likely to predict increased risk for engagement in relational aggression (Hemphill et al., 2010). Although most work on developmental trajectories of relational aggression has focused on early and middle childhood, some preliminary evidence indicates that relational aggression becomes somewhat less common in later adolescence. For instance, Cleverley, Szatmari, Vaillancourt, Boyle, and Lipman (2012) reported that most youth exhibited decreases in their engagement in indirect aggression across adolescence; however, a small subset of participants maintained fairly high levels of such conduct across the course of the study. Moreover, this group of youth contained approximately equal numbers of males and females. Karriker-Jaffe et al. (2008) reported that social aggression exhibited curvilinear change across adolescence, with a peak around 14 years of age (see also Karriker-Jaffe et al., 2013). Hemphill et al. (2012) found that levels of relational aggression decreased from grades 9 to 11 in a sample of Australian adolescents. Taken together, then, preliminary evidence suggests that relational aggression may increase in frequency from early childhood to early adolescence, peak in early to midadolescence, and then subsequently decline in frequency. Future prospective longitudinal studies are needed to confirm these developmental changes.

Developmental Manifestations of Relational Aggression The focus on mean change in overall levels of relational aggression, however, may obscure more fine-grained developmental differences in the manifestation of relationally aggressive behaviors. For instance, Crick et al. (2007) argued that relational aggression in young children

may most often be direct and easily observable (e.g., threatening to end a friendship; turning away from a peer to indicate exclusion; secret spreading in front of the victim). As children mature, however, they may become more sophisticated in their use of relational aggression; for instance, youth may be more likely to engage in indirect relationally aggressive behaviors or behaviors that involve third parties (e.g., rumor spreading; Crick et al., 2007). In fact, although sophisticated relationally aggressive behaviors have been observed in preschool children, these displays tend to be direct and easily observable by the targets (Ostrov et al., 2004; Fanger, Frankel, & Hazen, 2012). Longitudinal studies that explore developmental changes in the use of distinct relationally aggressive behaviors (e.g., threats of friendship withdrawal; spreading malicious gossip; social exclusion) may significantly advance our understanding of the developmental consequences of relational aggression. It is possible, for instance, that unsophisticated relationally aggressive behaviors (e.g., direct verbal threats to end a friendship) may be more strongly associated with maladjustment as compared to more covert behaviors (e.g., spreading gossip or engaging in exclusionary behaviors with a cover story; Fanger et al., 2012), particularly when enacted in developmental periods when such unsophisticated conduct is relatively atypical. Similarly, verbal and nonverbal relationally aggressive behaviors may have distinct implications for adjustment (see Blake, Kim, & Lease, 2011, for a recent application of this idea to social aggression). In addition, the contexts that elicit relational aggression may change across development. In young children, relational aggression is more likely to occur in sibling relationships than in friendships; however, over time, youth increasingly use these behaviors against friends (Stauffacher & DeHart, 2006), and, by middle childhood, friendships are an important context for the use of relational aggression (Crick & Grotpeter, 1996). With the emergence of romantic relationships in adolescence, youth may begin to use relational aggression against romantic partners (Linder, Crick, & Collins, 2002). In fact, some evidence indicates that in young adulthood, relational aggression more frequently occurs in the context of romantic relationships than in friendships, perhaps because these relationships become particularly salient during this developmental period (Goldstein, 2011; although see Goldstein & Tisak, 2004). Moreover, youth may target others' romantic relationships as a means of social manipulation or control (e.g., stealing a friend's boyfriend; Crick et al., 2007). However, much of the research regarding relational aggression has focused on the use of these behaviors in same-sex peer groups, and most measures fail to assess the relational context of aggression (Goldstein & Tisak, 2004). An important question for future research is whether there is continuity across relationships (e.g., friends, romantic relationships) in the use of relational aggression. Longitudinal studies may provide important insights regarding the potential transfer of relationally aggressive behaviors across developmental contexts (e.g., from sibling relationships to friendships; Stauffacher & DeHart, 2006).

Developmental Challenges in Assessment A number of assessment techniques are available for the measurement of relational aggression across developmental periods, including systematic observational procedures in the classroom and on the playground (e.g., McNeilly-Choque, Hart, Robinson, Nelson, & Olson, 1996;

Ostrov & Keating, 2004), semistructured observations (Ostrov et al., 2004), observational ratings (Murray-Close & Ostrov, 2009), peer nomination techniques (Crick & Grotpeter, 1995; McNeilly-Choque et al., 1996), teacher reports (Crick, 1996; McNeilly-Choque et al., 1996), structured interviews (Tackett & Ostrov, 2010), parent reports (Casas et al., 2006; Ostrov & Bishop, 2008), and self-reports (Little et al., 2003; Murray-Close, Ostrov, Nelson, Crick, & Coccaro, 2010). Details regarding methods for measuring relational aggression have been discussed elsewhere (Crick et al., 2007); however, two methodological challenges to the assessment of relational aggression merit comment because they have significant implications for the study of developmental change in relational aggression. First, there is currently a lack of clarity in the literature regarding items that assess relational, rather than social or indirect, aggression. As discussed, there is considerable conceptual overlap between these forms of aggression (Archer & Coyne, 2005), and it is not uncommon for researchers to use measures developed to assess relational aggression (or slight adaptations of these measures) in studies of indirect or social aggression. Given the similarities among social, indirect, and relational aggression (Archer & Coyne, 2005), this approach is understandable and in many circumstances may not have important implications for study findings. However, distinctions between these behaviors may be important for some critical developmental questions. For instance, the transition to indirect aggressive behaviors hypothesized to occur as youth get older (Björkqvist et al., 1992; Crick et al., 2007) underscores the possibility that developmental differences in aggressive behavior may partially depend on the subtype of aggression being assessed. More specifically, relational aggression may be more prevalent than indirect aggression among young children because relationally aggressive behaviors are often very direct (e.g., covering one's ears to ignore a peer). Additionally, the processes that underlie developmental increases in relational aggression may differ from those that give rise to increases in social or indirect aggression. For instance, as damage to relationships is the defining feature of relational forms of aggression, these behaviors may be most strongly tied to increases in close and intimate relationships (see Murray-Close et al., 2007) and most predictive of adjustment problems related to impairment in close relationships. We believe that it will be important for researchers to develop instruments that can better distinguish between these related, but conceptually distinct, subtypes of nonphysical aggression. In an example of efforts in this area, Nelson, Springer, et al. (2008) developed a coding scheme to identify subtypes of aggression in emerging adults' descriptions of peerbased mean behaviors. This scheme, including direct and indirect relational aggression, ignoring, gestural nonverbal aggression, and passive aggression, could be adapted to assess subtypes of aggressive behavior in youth. Results from this study documented important differences in emerging adults' engagement in these behaviors and highlighted how measures that are nuanced and theoretically based may help elucidate the processes underlying the development of relational aggression. Second, efforts to characterize developmental change in relational aggression, as well as research on developmental differences in the correlates of relational aggression, are limited by methodological differences in the assessment of relational aggression across developmental

periods. Naturalistic systematic observations have primarily been used with preschool-age children (e.g., Ostrov & Keating, 2004), likely because the heightened reactivity to traditional observation techniques often displayed among older children serves as a threat to the validity of the measures. Additionally, teacher reports may be more valid for use during early and middle childhood, as compared to adolescence, because relational aggression may be relatively direct, unsophisticated, and easily observed by onlookers (Crick et al., 2007). During adolescence, in contrast, parents and teachers may not be privy to many instances of relational aggression. Given the covert nature of relationally aggressive behaviors, Crick et al. (2007) suggest that peer nomination procedures are likely the most valid assessments of such conduct during late childhood and adolescence, although self-report measures of relational aggression are common and are meaningfully associated with outcomes (e.g., Little et al., 2003). The differences in the techniques that are most developmentally appropriate (and most commonly employed) across age groups create challenges for developmental psychopathology researchers. Specifically, true developmental change in relational aggression (and the correlates of such conduct) may be obscured by different assessment techniques and/or changes in the validity of a particular approach across development. For instance, declines in parentreported relational aggression may reflect true developmental change in such conduct or may instead reflect improvements in adolescents' abilities at concealing these behaviors from their parents. A critical direction for future research, then, is the continued development and refinement of methods for assessing relational forms of aggression across developmental periods.

Aggression and Gender Theoretical Perspectives Regarding Aggressive Girls Research on relational aggression challenged the long-standing assumption in the field of developmental psychopathology that aggression was highly atypical in females during childhood. In one highly influential theoretical account of aggression in females, Keenan and Shaw (1997) suggested that norms against aggression in girls, and accompanying socialization experiences, led females to shift from predominantly externalizing to internalizing problems across development. Additionally, girls' earlier maturation was argued to serve as a buffer against the emergence of problem behaviors such as aggression (Keenan & Shaw, 1997). Although several researchers suggested that girls may engage in aggressive behaviors during adolescence (e.g., Moffitt & Caspi, 2001; Silverthorn & Frick, 1999), there was a consensus among most developmental theorists that such conduct was quite rare among females during childhood. Empirical support from decades of aggression research bolstered these arguments; for instance, Maccoby and Jacklin (1974) argued that gender differences in aggression were one of the rare true gender differences. In their seminal publication on the topic of relational aggression, however, Crick and Grotpeter (1995) argued that the common assumption among developmental psychopathology researchers

that girls did not exhibit significant difficulties such as aggression during childhood, termed the benign childhood hypothesis (Crick & Zahn-Waxler, 2003), was incorrect. Instead, Crick and colleagues suggested that these apparent gender differences resulted from a failure to assess aggressive behaviors that were salient for females, such as relational aggression. Theoretically, relational aggression was expected to be more common than physical aggression among females for several reasons. First, girls were hypothesized to employ relationally aggressive behaviors because such behaviors are consistent with the focus on interpersonal relationships and social functioning typical in female peer groups (e.g., Crick & Grotpeter, 1995; Crick & Zahn-Waxler, 2003). Socialization against physical aggression from significant others, such as parents and peers, was also posited to reduce girls' reliance on physical aggression and promote the use of more socially acceptable and covert forms of aggression, such as relational aggression (Crick et al., 2007). Björkqvist et al. (1992) suggested that the relatively advanced cognitive maturation among girls, hypothesized by others to buffer girls from engagement in aggressive conduct (Silverthorn & Frick, 1999), may actually facilitate the use of more sophisticated aggressive behaviors, such as indirect aggression. Furthermore, girls' lower physical strength was argued to limit their capacity to effectively use physical forms of aggression, increasing the appeal of more indirect approaches (Björkqvist, 1994).

Gender Differences in Relational Aggression Despite these theoretical claims, however, research findings regarding gender differences in relational aggression have been mixed. Findings from several early studies on relational aggression provided evidence that girls were more likely than boys to engage in such conduct from preschool through childhood (see Crick et al., 2007, for review). However, Card, Stucky, Sawalani, and Little (2008) provided meta-analytic findings indicating that although gender differences in indirect aggression favoring girls were statistically significant, they were too small to be meaningful. Nonetheless, the study of relational aggression remains critical in research efforts to understand aggressive behavior problems in girls. Females are relatively unlikely to engage in physical forms of aggression, and there appears to be greater overlap between physical and indirect aggression in males compared to females (Card et al., 2008). These findings suggest that aggressive boys may frequently employ both physical and indirect aggression, whereas aggressive girls may preferentially select indirect or relational forms of aggression; that is, the modal type of aggression (i.e., within-group gender difference) is relational rather than physical among girls (e.g., Ostrov, Kamper, Hart, Godleski, & BlakelyMcClure, 2014; Putallaz et al., 2007; see also Ostrov & Godleski, 2010). In fact, findings from recent person-centered analyses indicate that across several populations (i.e., community, voluntary residential treatment, and involuntary detained), adolescent girls are more likely than boys to be classified as highly relationally, but not physically, aggressive (Marsee et al., 2014). As a result, studies that do not assess relational, indirect, or social forms of aggression may fail to identify the vast majority of aggressive females. Additionally, several researchers have emphasized the importance of moving beyond a focus on mean-level differences in relational aggression to instead explore gender differences in developmental trajectories, mechanisms, and outcomes associated with such conduct (e.g.,

Ostrov & Godleski, 2010; Spieker et al., 2012; Underwood, 2003). In fact, several of the developmental processes thought to promote engagement in relational aggression (e.g., increases in intimacy in close friendships; pubertal changes) may differ between males and females, potentially leading to meaningful gender differences in trajectories of relational aggression over time (e.g., Kistner et al., 2010; Murray-Close et al., 2007). For instance, gender differences in pubertal timing may result in a relatively early increase in relational aggression among girls during the transition to adolescence (Kistner et al., 2010). Card et al. (2008) failed to find meta-analytic evidence that gender differences in indirect aggression differed across developmental periods; however, measurement differences across these periods may have obscured gender differences in true developmental change (Card et al., 2008), highlighting the need for longitudinal research in this area. In fact, some longitudinal research has provided preliminary evidence for distinct patterns of developmental change in relational aggression in males versus females. For instance, several studies have reported that girls were more likely than boys to exhibit increases in relational aggression (or to maintain relatively high levels of such conduct) as they approached adolescence (Kawabata et al., 2012; Kistner et al., 2010; Murray-Close et al., 2007; Spieker et al., 2012). However, other studies have failed to find gender differences (Cleverley et al., 2012; Karrifer-Jaffe et al., 2008; Underwood et al., 2009). Clearly, additional research exploring gender differences in longitudinal change in relational aggression, with attention on the methods used to assess such conduct, is needed. It will also be important for researchers to consider the question of gender differences across relational contexts and across subtypes of relationally aggressive conduct. In fact, some evidence indicates that relational aggression is more common than physical aggression in girls', but not boys', friendships (Crick & Nelson, 2002). Goldstein (2011) reported that although young adults did not differ in their use of relational aggression against friends, females were more likely than males to engage in such conduct in the context of romantic relationships (for additional research regarding gender differences in romantic relational aggression, see Goldstein, Chesir-Teran, & McFaul, 2008; Linder et al., 2002). These preliminary findings suggest that across developmental periods, gender differences may exist in some (e.g., romantic relationships) but not other (e.g., friendships) relational contexts. In addition, research in this area may benefit from a more fine-grained analysis of the types of relationally aggressive behaviors used against relationship partners. For instance, Nelson, Springer, et al. (2008) found that college students reported that males most often engaged in direct physical aggression (e.g., hitting), particularly against other males. In contrast, participants reported that females frequently engaged in indirect relational aggression (e.g., gossip) against both males and females and in direct relational aggression (e.g., guilt trips) against males. These findings suggest that gender differences in relational aggression may depend on the gender of the target as well as whether the relational aggression is direct or indirect in nature. Future work is necessary to elucidate whether gender differences exist across relational contexts (e.g., friendships, romantic relationships) and relationally aggressive behaviors (e.g., indirect, direct). It will also be important to assess whether the magnitude of these gender differences changes across development.

Risk Factors: Biobehavioral Processes Genetic Risk Developmental psychopathologists have highlighted the need for interdisciplinary research, including neurobiological, neuropsychological, and psychophysiological approaches, in the study of the development of maladaptive behavior patterns for decades (e.g., Cicchetti, 1993). Several biobehavioral perspectives regarding the development of relational aggression have emerged in recent years, including studies of genetic risk for involvement in relationally aggressive behaviors. Using a quantitative genetic design, Brendgen et al. (2005) examined the relative genetic and environmental contributions to physical and social aggression in a sample of 6-year-old twin pairs. Findings indicated that social aggression was only weakly related to genetic factors (20% of the variance for teacher-reported social aggression; 23% of the variance for peer-reported social aggression) and was associated with both shared and nonshared environmental influences. These findings contrasted with the results for physical aggression, which indicated substantial heritability (63% of the variance for teacher-reported physical aggression; 54% of the variance for peer-reported physical aggression) and virtually no influence of shared environmental factors. Interestingly, although the genetic influence on social aggression was relatively low, it overlapped with genetic risk for physical aggression, suggesting similar genetic etiological factors for both forms of aggression (Brendgen et al., 2005). In contrast, the nonshared environmental risk factors for physical versus social aggression appeared to be relatively distinct. In their study of 6- to 18-year-old twins, Tackett, Waldman, and Lahey (2009) reported relatively high estimates of genetic influence on a latent variable including mother- and selfreported relational aggression (63% of the variance). However, as in the work of Brendgen et al. (2005), shared environmental risk emerged; the authors suggested that relational aggression may be more susceptible than physical aggression to parent and shared peer group influences (Tackett et al., 2009), highlighting the important role of environmental factors in the emergence of such conduct. The role of the shared environment in the development of relational aggression has not always been replicated in genetic studies, however. In a follow-up study using a subsample of participants from Brendgen et al. (2005) when youth were 7 years old, Brendgen et al. (2008) failed to find evidence for shared environmental influences on a composite of teacher- and peer-reported social aggression. Tackett et al. suggested that distinct informants may account for some of the mixed findings in this area and highlighted the need for greater empirical attention to measurement of relational aggression in behavior genetic designs. Behavioral genetics studies provide significant insight regarding potential etiological factors involved in the development of relationally aggressive conduct. However, a number of significant questions remain. For instance, emerging molecular genetic studies have documented specific candidate genes in the development of antisocial behavior, particularly in the context of environmental risk (e.g., Kim-Cohen et al., 2006). Given evidence of similar genetic risk for both forms of aggression in quantitative genetic designs (Brendgen et al.,

2005), researchers should extend this work to molecular genetic studies to explore whether similar genetic polymorphisms are related to physical and relational forms of aggression. It will also be essential to evaluate whether similar environmental stressors are involved in gene-environment (G×E) interactions in the development of physical versus relational aggression. Brendgen et al. (2008) reported a G×E interaction in the association between friends' physical aggression and participants' own involvement in such conduct; a parallel interaction was not found in participants' social aggression. These findings highlight the possibility that distinct environmental stressors may interact with genetic risk in predicting relationally aggressive behaviors, as compared to more traditionally studied forms of aggression (e.g., physical aggression). Finally, genetically informed, longitudinal designs are necessary to evaluate the extent to which genetic factors contribute to stability and change in relational aggression over time and whether the patterns of these contributions are similar for physical and relational forms of aggression (see Neiderhiser, Reiss, & Hetherington, 1996, for an application to antisocial behavior).

Psychophysiological Stress Systems In addition to genetic risk, researchers have recently begun to examine the role of individual differences in functioning of psychophysiological stress systems, including the autonomic nervous system (ANS; see Murray-Close, 2013a), in the development of relational aggression. The ANS is comprised of two distinct branches, the sympathetic nervous system (SNS), which coordinates the body's fight-or-flight responses to stress, and the parasympathetic nervous system (PNS), which reflects rest-and-digest functions. In several theoretical formulations, low SNS arousal is hypothesized to serve as a risk factor for the development of aggressive conduct. According to fearlessness theory, low SNS activity reflects temperamental fearlessness (Ortiz & Raine, 2004), which in turn may interfere with the efficacy of socialization against aggressive behavior. Fearless youth may be relatively unresponsive to punishment and unconcerned about potential negative repercussions of engaging in aggressive conduct (e.g., retaliation from victims). From a stimulation-seeking perspective, low SNS is experienced as an aversive physiological state; thus, youth may engage in aggressive behaviors in an effort to increase their arousal to more comfortable levels (Sijtsema, Veenstra, et al., 2010). However, some researchers have argued that heightened, rather than blunted, SNS arousal is a risk factor for aggressive conduct, particularly when SNS overarousal occurs following aversive or threatening events (e.g., Hubbard et al., 2002). From this perspective, exaggerated SNS reactivity to negative events is hypothesized to reflect dysregulated emotional reactions to stress (e.g., anger), thus increasing risk for aggressive responses (see Murray-Close, 2013a). Indices of PNS functioning, in contrast, are argued to reflect biological differences in regulatory abilities, such that high resting PNS arousal, and greater PNS withdrawal following challenge, portend greater emotional and attention regulation capacities (Calkins & Keane, 2004; Porges, 2007). This greater self-regulation, in turn, is expected to facilitate the inhibition of aggressive impulses. To date, the vast majority of research regarding the association between ANS arousal and aggression has focused on physical forms of aggressive behavior (see Murray-Close, 2013a,

for review). However, an emerging body of research explores the association between ANS arousal and relational forms of aggression, with some work documenting distinct correlates of relational, as compared to physical, aggression (Murray-Close & Crick, 2007). Some of this research has provided evidence that exaggerated stress reactivity is related to increased risk for relational aggression. In early work in this area, for instance, Murray-Close and Crick (2007) found that heightened systolic blood pressure reactivity was associated with relational aggression in girls but not boys. More recently, in a sample of female emerging adults, MurrayClose (2011) found that heightened SNS reactivity to stress was associated with relational aggression against romantic partners, particularly among women who also exhibited additional risk factors for aggression (e.g., hostile attribution bias for romantic relational provocations). However, consistent with fearlessness and/or stimulation-seeking theories, other research has documented blunted SNS arousal among relationally aggressive youth. For instance, Sijtsema, Shoulberg, and Murray-Close (2011) reported that blunted SNS reactivity was associated with relational aggression in a sample of female children and adolescents. In a study with preschoolers, Gower and Crick (2011) found that low resting heart rate and blood pressure were associated with elevated levels of relational aggression; in some analyses, findings emerged primarily in youth who also exhibited poor effortful control. Although more limited, some studies have also documented associations between PNS withdrawal and relational aggression, although the direction of effects has varied across studies (e.g., Murray-Close, 2011; Sijtsema et al., 2011). Preliminary research regarding the association between relational aggression and neuroendocrine indices of stress, such as cortisol, has also provided support for the notion that underarousal of stress systems serves as a risk factor for relationally aggressive conduct. Cortisol, a hormone released by the adrenal cortex in response to stressful experiences, exhibits a diurnal rhythm with a peak in cortisol levels 30 minutes after waking followed by gradual decrease throughout the day. Murray-Close, Han, Cicchetti, Crick, and Rogosch (2008) found that relational aggression was associated with low morning cortisol and blunted change in cortisol across the day. However, findings are preliminary, and some researchers have failed to find associations between cortisol and indirect aggression (e.g., Vaillancourt & Suderani, 2011). The mixed findings regarding the associations between both ANS and neuroendocrine stress systems and relational aggression highlight the possibility that both heightened and blunted stress arousal may predict elevated levels of relational aggression, depending on additional moderating factors. For instance, the physiological stress profiles associated with relational aggression may vary for males versus females (Murray-Close et al., 2014; see Murray-Close, 2013a, for review). Effects may also depend on the presence of additional risk factors (e.g., relational victimization; Murray-Close, 2011; Murray-Close et al., 2014; see also MurrayClose, Holland, & Roisman, 2012) or whether the aggression is proactive (i.e., planful, unemotional, and goal-directed) or reactive (i.e., emotional, retaliatory) in function. In fact, Murray-Close and Rellini (2012) recently reported that blunted ANS reactivity to stress was associated with proactive relational aggression whereas exaggerated ANS reactivity was associated with reactive relational aggression; however, findings emerged only among women

with a history of sexual abuse. Interestingly, the moderating role of contextual risk has varied across studies, with some research indicating that physiological arousal is more strongly related to relational aggression among youth without high levels of contextual risk (e.g., without a history of abuse; Murray-Close et al., 2008; see the “Multilevel Perspectives” section for a discussion of potential explanations regarding these differing findings). A relatively understudied moderator in research regarding the association between relational aggression and physiological reactivity is the type of task used to elicit stress reactivity; in fact, Murray-Close and colleagues (2014) recently documented that girls' perpetration of relational aggression was at times associated with distinct profiles of reactivity to relational (e.g., experiences of exclusion) versus instrumental (e.g., property damage) peer stressors. It will be essential for future research to consider moderators of the association between stress physiology and relational aggression, with particular attention to the processes thought to underlie these associations.

Additional Biological Factors Several potential biological risk factors for relational aggression should be incorporated in future research in this area. For instance, despite the large body of research documenting associations between androgens such as testosterone and physically aggressive behavior (Archer, Graham-Kevan, & Davies, 2005), limited research has examined whether androgens are implicated in relational aggression. In preliminary research in this area, testosterone was positively associated with relational aggression in a sample of 9-year-old children, once effects of gender were controlled (Sánchez-Martín et al., 2011). Coyne, Manning, Ringer, and Bailey (2007) found that low right-left ratios of the length of the index finger to ring finger, thought to reflect greater exposure to prenatal androgens, were related to indirect aggression in a sample of women. Taken together, these results highlight the possibility of both organizational and activational effects of androgens in the development of relational aggression. Emerging research has also documented associations between event-related potentials and relational aggression, perhaps reflecting attentional deficits among aggressors (Godleski, Ostrov, Houston, & Schlienz, 2010). Additional psychophysiological indices thought to reflect motivational and emotional processes, such as startle and electroencephalography, have been found to be related to physical aggression (e.g., Peterson, Shackman, & Harmon-Jones, 2008; Verona & Kilmer, 2007). It will be important for researchers to extend this work to include relational forms of aggression, particularly in research areas that have documented gender differences in the prediction of physically aggressive conduct (e.g., Verona & Kilmer, 2007). Finally, researchers should incorporate social affective neuroscience perspectives to explore the neural mechanisms underlying relationally aggressive behaviors, including investigating whether neuroscience models of physical aggression (e.g., Davidson, Putnam, & Larson, 2000) can be applied to relational forms of such conduct. In one recent functional magnetic resonance imaging study with adolescent girls, experiences of relational victimization were related to decreased activity in brain regions associated with executive functioning when viewing peer faces with varying affective expressions (Baird, Silver, & Veague, 2010). However, to our

knowledge, researchers have not yet examined the neural activation patterns associated with the perpetration of relational aggression using functional magnetic resonance imaging methods, despite the central role ascribed to emotional processes in such conduct (see the “Emotion and Emotion Regulation” section). Given the hypothesized importance of reactivity to relational stressors among relationally aggressive youth (Crick, Grotpeter, & Bigbee, 2002), these studies may benefit from incorporating procedures that can assess neural activation patterns associated with social exclusion (e.g., Eisenberger, Lieberman, & Williams, 2003).

Risk Factors: Cognitive and Emotional Processes The Crick and Dodge (1994) Reformulated Social Information Processing Model of Children's Adjustment It has been over two decades since Crick and Dodge (1994) published their seminal review and reformulation of the social information processing (SIP) model of children's adjustment (originally proposed by Dodge, 1986). This review provided clear evidence that unique tendencies in processing cues from the social world could help explain behavioral differences between children. In particular, the accumulated evidence strongly indicated that heightened levels of physical (or overt) aggression were consistently associated with processing biases and deficits at every step of the model. Experimental and intervention designs have been particularly effective in demonstrating a causal role of these SIP mechanisms in children's engagement in aggressive acts. The proposition of the model (see Figure 13.1 for a representation) is that children come to process social cues in a consistent, automated fashion which can predispose them to aggressive action against others (whether warranted or not). The steps of the reformulated model, leading up to behavioral enactment (step 6), include (1) encoding of external and internal cues, (2) mental representation and interpretation of those cues, (3) clarification or selection of a social goal, in the context of current interpretation, (4) access or construction of potential responses, and (5) response decision (which response seems most appropriate, given current interpretation and social goals). Prior experience (in the database) provides context for current processing.

Figure 13.1 The Reformulated Social Information Processing Model. Source: Crick, N. R., & Dodge, K. A. (1994). A review and reformulation of social information-processing mechanisms in children's social adjustment. Psychological Bulletin, 115, 74–101. Copyright © 1994 by the American Psychological Association. Reprinted with permission.

During the first two steps of this model, individual differences in children's processing are evident as they selectively attend to cues, which are then subject to interpretation. Aggressive children are more likely to attend to cues consistent with negative expectations of the intent of others (based on prior experiences and associated social schemata; Dodge & Tomlin, 1987). The aggressive child's enhanced attention to cues which may be perceived as hostile (Gouze, 1987) facilitates interpretation of those cues in a biased manner. In particular, aggressive children tend to have a hostile attributional bias (HAB), in which they are more likely to perceive hostile intent, even when cues are ambiguous and therefore do not lend themselves to such an interpretation (Crick & Dodge, 1994). This relationship between HAB and social adjustment tends to be quite robust across studies, at least in regard to physical aggression (Orobio de Castro, Veerman, Koops, Bosch, & Monshouwer, 2002); in effect, aggression

serves as a defense against perceived provocation. At step 3, children select a goal or desired outcome for the current situation. These goals may either enhance or detract from optimal relationship functioning (Slaby & Guerra, 1988). In this regard, aggressive children are more likely to embrace goals that damage relationships (e.g., revenge or instrumentality of aggression) as they navigate social interactions, consistent with their tendency to perceive hostile cues from others. At step 4, children are hypothesized to access possible responses from memory or to construct novel responses in unfamiliar social situations. The number of responses, their order, and the tenor of each individual response may be informative in the evaluation of the child's processing at this step. The response repertoires of aggressive children are characterized by more aggressive and less prosocial approaches (Quiggle, Garber, Panak, & Dodge, 1992). Finally, at steps 5 and 6, the potential responses are evaluated and the most appropriate response selected for behavioral enactment. Aggressive children tend to evaluate aggressive responses more positively and are less likely to believe that competent responses will be effective in dealing with peers (Quiggle et al., 1992). These evaluations appear to be particularly relevant to engagement in proactive aggression (Crick & Dodge, 1996). Finally, at step 6, the behavior is displayed. Importantly, consistent with the model, the manner in which peers respond to this behavior will impact subsequent encoding and processing of cues. That is, children that enact an aggressive behavior will often experience retaliation from peers, which may confirm their prior intent attributions and allow for faster automated processing in similar social situations in the future (Crick & Dodge, 1994). Gender, Subtypes of Aggression, and Hostile Attribution Bias At the time of the Crick and Dodge (1994) review of SIP research, physical and verbal aggression had been the focus of prior work, largely because constructs such as indirect and relational aggression were just beginning to capture the attention of developmental scientists. There was promise evident, however, in the application of the SIP model to relational aggression. For instance, Crick (1995) hypothesized that relationally aggressive children, like their physically aggressive counterparts, would also exhibit HAB. However, Crick argued that relational disputes (e.g., provocations with a relational tone, such as being left out) would be particularly provocative for relationally aggressive children, especially girls, because these youth view such behavior as highly distressing and disruptive to their interpersonal goals. In contrast, Crick argued that boys would find instrumental disputes (provocations involving physical dominance, territory, and material objects) to be most distressing and thus would react with physical aggression. This hypothesis raised the possibility that HAB may be context-specific; that is, relational scenarios were not expected to be provocative to physically aggressive children, and instrumental scenarios were not expected to be provocative to relationally aggressive children. In the first two studies that assessed the validity of these hypotheses (Crick, 1995; Crick et al., 2002), results generally aligned with expectations. Peer-nominated relational aggression corresponded with higher levels of children's self-reported relational HAB and emotional

distress. Girls were also more likely than boys to feel such distress. Several recent studies have provided additional support for Crick's (1995) initial hypotheses. For instance, Gentile, Coyne, and Walsh (2011) found instrumental HAB mediated the association between fall media violence exposure and spring physical aggression, particularly for boys. In addition, relational HAB mediated the relation between fall media violence exposure and spring relational aggression, and this effect was stronger for girls. With a sample of preadolescents, Yeung and Leadbeater (2007) found that relational HAB was associated with relational aggression concurrently as well as over time. Moreover, HAB for relational provocations partially mediated the association between relational victimization and concurrent relational aggression, though only at Time 1 (see also Ostrov, Hart, Kamper, & Godleski, 2011; Ostrov & Godleski, 2013). In a sample of adults, Murray-Close et al. (2010) found that HAB for relational provocations and feelings of distress were associated with relational aggression against romantic partners, controlling for relational aggression against peers. However, others have reported findings that do not confirm the expected conceptual specificity linking subtypes of aggression to forms of HAB. Godleski and Ostrov (2010) explored temporal associations between HAB and aggression, using data from the National Institute of Child Health and Human Development Study of Early Child Care and Youth Development (NICHD SECCYD). In longitudinal analyses, third-grade aggression, gender, and relational and instrumental HABs (composites based on yearly assessments from third to fifth grade) were entered as predictors of sixth-grade subtypes of aggression. When controlling for earlier aggression and gender, instrumental HAB rather than relational HAB predicted later relational aggression and physical aggression. When temporal order was reversed (third-grade aggression predicting the HAB composites), both relational and physical aggression were associated with relational and instrumental HABs. Taken together, then, the findings suggested a more generalized association between aggression and HAB. Additional studies have failed to find any association between HAB for relational provocations and relational aggression. For example, Crain, Finch, and Foster (2005) reported that relational HAB showed no connection with relational aggression in one sample of girls, and, in the second female sample, HAB for relational provocations was associated with lower levels of relational aggression. In an effort to further validate the Crick hypothetical scenarios for measuring HAB, Nelson, Mitchell, and Yang (2008) utilized confirmatory factor analysis to test whether responses to the different provocation contexts (instrumental and relational) could be statistically distinguished and whether all of Crick's original scenarios would appropriately load with the expected latent factor (two additional items were included in data collection as backup). Results showed that 9 of 10 Crick scenarios loaded on the appropriate factor; one relational provocation was replaced. The resulting instrumental HAB factor was related to physical aggression for boys, but children's relational HAB was not associated with relational aggression. In addition, the authors reported that for three of the five relational provocations, the majority of children in the study (58%–70%) responded with HAB. Accordingly, most relational scenarios were highly provocative and did not appear to elicit HAB from a relationally aggressive minority alone.

Moderators and Methodological Issues in the Hostile Attribution BiasAggression Connection These mixed findings highlight the possibility that there may be important moderators of the association between HAB and relational aggression. For example, Mathieson and colleagues (2011) recently examined moderators suggested by Crick's relational vulnerability model. They found relational HAB to predict relational aggression only for girls who experienced high levels of relational victimization by peers and also reported high emotional distress in reaction to Crick's relational provocation scenarios. This study has clear empirical and clinical implications, suggesting that relational factors may cluster together to increase risk for engagement in relational aggression, particularly among girls. In addition, proactive and reactive functions of relational aggression might be differentially associated with HAB. Prior research regarding physical aggression suggests that HAB is more strongly associated with reactive rather than proactive functions (Crick & Dodge, 1996). Although limited research has been conducted in the area of relational aggression, preliminary findings support the suggestion that HAB for relational provocations may be more strongly associated with reactive than proactive functions of relational aggression. For instance, Bailey and Ostrov (2008) found that reactive physical aggression corresponded uniquely with instrumental HAB, whereas reactive relational aggression was uniquely associated with relational HAB. In a study with adults, Murray-Close et al. (2010) reported that HAB for relational provocations was uniquely associated with reactive but not proactive relational aggression. In addition to moderators, the mixed findings may be partly due to methodological, analytical, and sample differences that limit the direct comparison of findings. For instance, Crain et al. (2005) assessed HAB in response to ambiguous situations in friendships rather than the mix of peer and friend situations captured in Crick's original measure. Additionally, their hypothetical situations focused on female characters only. Given Crick's (1995) suggestion that the link between HAB and aggression may be context-specific, these findings underscore the possibility that HAB may be more strongly associated with relational aggression in some contexts (e.g., in the larger peer group) than others (e.g., in close friendships). In physical aggression research, intent attributions have been assessed in many different ways (e.g., hypothetical scenarios, videotaped scenarios with actors), and effect sizes vary across these methodologies (Orobio de Castro et al., 2002). Relational provocation assessments have predominantly focused on hypothetical scenarios, so there is much work yet to be done to incorporate and compare results of different methodologies. The studies have also varied in that some incorporated a categorical approach to defining aggression (i.e., using cutoffs to establish extreme groups for aggression; e.g., Crick, 1995; Crick et al., 2002) whereas others have utilized a dimensional approach (i.e., using continuous aggression scores; e.g., Crain et al., 2005; Nelson, Mitchell, et al., 2008). Only Godleski and Ostrov (2010) have compared the utility of a categorical and a dimensional approach with the same sample. Research is also needed to assess whether these processes might differ across samples. Preliminary research examining the association between HAB and relational aggression in

clinical populations has provided mixed results. MacBrayer, Milich, and Hundley (2003), for example, found that children with a disruptive behavior disorder (attentiondeficit/hyperactivity disorder [ADHD], oppositional defiant disorder, bipolar) were more likely to exhibit HABs of both types. In contrast, Mikami, Lee, Hinshaw, and Mullin (2008) did not find girls with ADHD to have higher HAB scores. Moreover, connections between SIP and aggression were actually stronger for comparison girls than for girls with ADHD. Relational provocation scenarios may also be more or less provocative across different gender, cultural, and age groups, which may in turn enhance or weaken the strength of the association between HAB and aggression. Social Information Processing and Relational Aggression: Beyond Hostile Attribution Bias Although the majority of research regarding SIP processes and relational aggression has focused on HAB, several studies have also considered other steps in the SIP model as well as the presence of latent knowledge structures in the database (which informs each step of the model). For example, Arsenault and Foster (2012) recently examined how attention and memory processes might be biased in aggressive children, leading to difficulties at step 1 of the SIP model—encoding of cues. Children saw videos of unambiguous overtly and relationally aggressive provocations. During the videos, a red light would flash near the video screen, and the child had to press the space bar at that time. Reaction times were measured to assess children's latencies as an index of attention. Children's accuracy of recall to video content was also assessed. Results showed that relational aggression was associated with attention shifting and free recall for the relationally aggressive provocations, even after controlling for overt aggression. Relationally aggressive children, therefore, were particularly fixated on relationally aggressive events shown on the videotape. Crick and Werner (1998) examined response decision, the fifth step of the model, in the prediction of relational and physical (overt) aggression in third- to sixth-grade children. Children responded to hypothetical provocation scenarios (both relational and instrumental) in which the harmful intent of the provocateur was clear. They rated the appropriateness of physically aggressive, relationally aggressive, and prosocial responses to the provocation. Results showed that physically aggressive boys and girls positively evaluated physically aggressive responses to instrumental conflicts. In contrast, relationally aggressive children were not more likely to positively evaluate relationally aggressive responses to relational provocations. Contrary to expectations, physically aggressive girls saw physically aggressive responses more positively in the context of a relational provocation. Moreover, relationally aggressive boys favored relationally aggressive responses to instrumental provocation. Gender differences showed that boys were more positive about physically aggressive responses, and girls favored relationally aggressive responses. Some of these results are consistent with the original conceptualization of Crick (1995), whereas others differ. Finally, several studies have shown that normative beliefs about the appropriateness of aggression are related to engagement in relational aggression, potentially because these beliefs enhance selection of aggressive responses at step 5 of Crick and Dodge's (1994) model. For

instance, Poteat, Kimmel, and Wilchins (2011) found that beliefs regarding the appropriateness and effectiveness of violence were associated with bullying, physical fighting, and relational aggression for both male and female adolescents (see also Leff et al., 2014). Goldstein and Tisak (2010) found that adolescents viewed gossip and physical aggression as very wrong, but social exclusion was considered somewhat acceptable. Moreover, the more acceptable study participants presumed physical aggression or gossip to be, the more likely they were to admit engaging in physical aggression and relational aggression, respectively (no findings emerged for social exclusion). Werner and Hill (2010) assessed whether children's individual normative beliefs and classroom-level norms favoring relational aggression would predict increased engagement in such conduct over the course of one year. Results showed that approval of relational aggression increased with the transition to middle school. Moreover, increases in relational aggression were predicted by individual norms favoring relational aggression and a peer group context highly supportive of relational aggression.

Toward an Integrated Gender-Linked Model of Aggression Recently an integrated model for understanding gender differences in the development of physical and relational aggression in early and middle childhood was proposed by Ostrov and Godleski (2010). This model, depicted in Figure 13.2, posited a new framework that expanded on the existing social-cognitive and gender schema models and specifically integrated the SIP model of children's adjustment (Crick & Dodge, 1994) with the Schematic-Processing Model of Sex Role Stereotyping (Martin & Halverson, 1981). The model begins with the traditional six steps of the SIP model. However, Ostrov and Godleski proposed that the gender schematic processing model can be fully incorporated in the SIP model via the database and that these gendered processing patterns influence each of the traditional six SIP steps.

Figure 13.2 Gender-Linked Model of Aggression Subtypes. Source: Ostrov, J. M., & Godleski, S. A. (2010). Toward an integrated gender-linked model of aggression subtypes in early and middle childhood. Psychological Review, 117, 233–242. Copyright © 2010 by the American Psychological Association. Reprinted with permission.

At step 1 of the SIP model, Ostrov and Godleski (2010) argued that past gender-relevant peer experiences, stored as memories, would be accessed during ambiguous provocation situations when encoding cues. Gender schemas were argued to influence the interpretation of cues at step 2 and the clarification of goals at step 3. At step 4, when generating possible responses to the provocation, behavioral responses that are most consistent with children's gender identity, self-construal, and past experiences (i.e., influenced by various socialization agents and the individual's history of adaptation) are more likely to be activated and accessible. Next, the authors proposed that youth will be probabilistically more likely to engage in aggressive behaviors that they consider self-relevant and typical of their gender. Finally, after enacting an aggressive behavior, youth will be more prone to remember such conduct when it is deemed gender- and self-relevant than when the behavior is inconsistent with their gender- and selfschemas. This gender-linked model of aggression posits several predictions. For example, this model suggests within-group gender differences and preferences. That is, children are theorized to prefer to use gender-consistent aggressive behaviors and avoid the display of gender-

inconsistent aggressive acts. From this perspective, the critical question regarding gender differences in relational aggression is not whether girls are more relationally aggressive than boys. Instead, this model posits that girls will more frequently select relationally aggressive behaviors than physically aggressive behaviors whereas boys will more frequently select physically aggressive behaviors than relationally aggressive behaviors. In fact, this prediction has been supported in prior research with a sample of preschoolers (Ostrov et al., 2014) and a large sample of ethnically diverse girls and boys in middle childhood (Putallaz et al., 2007). This model also predicts that gender will serve as an important moderator of associations between aggression and future social-psychological adjustment problems. This gender-linked model of aggression subtypes argues that because girls are more sensitive to close interpersonal relationships and base their aggressive reactions on interpersonally related social cognitions (e.g., HAB for relational provocations), they will exhibit more negative adjustment outcomes when they display and receive relational aggression from their peers (Ostrov & Godleski, 2010). Using the NICHD SECCYD, Spieker et al. (2012) found some support for this hypothesis. Specifically, initial levels of relational aggression predicted adolescent self-reports of depression for girls but not boys. In another study, relational aggression was significantly and uniquely associated with future relational victimization among preschool girls but not among preschool boys (Ostrov, 2008). Overall, this model suggests that the field should move toward more sophisticated questions that address causal mechanisms and developmental trajectories to more appropriately understand the role of gender in the development of aggression among children.

Emotion and Emotion Regulation A significant body of research has highlighted the important role of emotions in the development of aggressive behaviors. Frick and Morris (2004) identified two temperamental pathways to aggressive conduct, each characterized by a distinct set of underlying emotion processes. In the first pathway, dysregulated negative emotional experiences, particularly with respect to anger and frustration, promote antisocial and aggressive behavior. From this perspective, angry reactions to peer stress may increase the likelihood that a child will respond to real or perceived provocation with aggressive conduct. Exaggerated negative emotional responses to stress may also interfere with the development of the social-cognitive and social skills necessary to inhibit aggressive behaviors. In fact, emotion regulation deficits and emotion-related processes have been implicated as potential contributors and mediators in numerous models of the development of aggression (e.g., Shields & Cicchetti, 1998). Emerging research has documented an important role of negative emotionality and/or poor regulation of negative affect in relationally aggressive conduct. For instance, Ostrov et al. (2013) recently reported that anger was associated with increases in relational aggression in a sample of preschool children followed across 4 months. In work with child and adolescent samples, relational aggression has been found to be related to negative affect and neuroticism (Tackett, Kushner, Herzhoff, Smack, & Reardon, 2014), angry responses to provocation (Marsee & Frick, 2007), the inability to tolerate frustration and anger (Little et al., 2003; Musher-Eizenman et al., 2004), affective lability (Tackett et al., 2014), elevated levels of

hostility (Little et al., 2003), ineffective coping (including the tendency to be overwhelmed by emotions; Tackett et al., 2014), and poor ability to manage emotions during stress (Gower et al., 2014). Indices of anger and hostility are also associated with relational aggression against romantic partners in adulthood (Murray-Close et al., 2010), suggesting that these emotional processes may be important in the emergence of relational aggression across relational contexts. More general emotion regulation capacities, such as the ability to express emotion appropriately, have not always been found to predict relational aggression (e.g., Ostrov et al., 2013), potentially highlighting the significance of emotional regulation capacities specifically related to anger or frustration. In the second pathway to aggression identified by Frick and Morris (2004), relatively low levels of emotional arousal may serve as a risk factor for aggressive conduct. For this pathway, a failure to experience normative levels of fear is hypothesized to interfere with experiences of guilt and anxiety regarding transgressions; as a result, socialization efforts against aggression may be relatively ineffective among these youth. In fact, a large body of research suggests that low levels of fear, guilt, and empathy are risk factors for aggressive or antisocial conduct. For instance, psychopathy, a personality profile including impaired affective (e.g., unemotional style) and interpersonal (e.g., manipulativeness) processes, is related to severe antisocial conduct (Frick & Morris, 2004). In recent years, researchers have documented that emotional processes consistent with psychopathy (e.g., cold affect) are associated with indirect aggression (Coyne & Thomas, 2008; Czar, Dahlen, Bullock, & Nicholson, 2011) and romantic relational aggression (Coyne, Nelson, Graham-Kevan, Keister, & Grant, 2010; Czar et al., 2011) in college student samples. Although psychopathy is multifaceted and includes factors beyond emotional functioning, some evidence indicates that primary psychopathy (i.e., cold affect and social manipulativeness) is more strongly related to indirect aggression than secondary psychopathy (i.e., criminal tendencies, erratic lifestyle, impulsivity; Coyne & Thomas, 2008; Vaillancourt & Sunderani, 2011). Moreover, indirect and relational forms of aggression are specifically associated with the coldheartedness factor of psychopathy (Warren & Clarbour, 2009) and are related to low levels of empathy (Smits, Doumen, Luyckx, Duriez, & Goossens, 2011). Taken together, these findings suggest that low emotional arousal may play an important role in relational forms of aggression during early adulthood. Researchers have also reported relatively low levels of emotional arousal in relationally aggressive children and adolescents. In early research in this area, Kaukiainen et al. (1999) found that low empathy was related to heightened levels of indirect aggression in a sample of adolescents. More recently, Tackett and colleagues (2014) documented that a lack of empathy was associated with relational aggression in a large sample of adolescents. Similarly, MacEvoy and Leff (2012) reported that lower levels of sympathy for victimized peers were associated with more relational aggression in a predominantly African American sample of third to fifth graders. Longitudinal work highlights the potential role of emotional underarousal in the emergence of relational aggression; for instance, Batanova and Loukas (2011) found that empathic concern was related to decreases in relational aggression in middle school students followed across 1 calendar year.

In related work, several researchers have documented that callous-unemotional (CU) traits, characterized by low levels of guilt and empathy, are associated with relational aggression (e.g., Lau & Marsee, 2013). For instance, Marsee and Frick (2007) reported that CU traits were positively associated with proactive relational aggression in a sample of incarcerated female adolescents. Similarly, Kerig and Stellwagen (2010) found that CU traits were positively correlated with relational aggression during early adolescence, although this effect did not persist once impulsivity was controlled. In a study with Japanese children and adolescents, Onishi, Kawabata, Kurokawa, and Yoshida (2012) reported that individuals who experienced feelings of guilt for engaging in relational aggression were less likely to choose to participate in such conduct in response to a series of hypothetical scenarios. Overall, then, findings across developmental periods have provided support for a relatively unemotional pathway to relational aggression. Several factors may moderate the association between emotional processes and relational aggression. For instance, as reactive functions of aggression reflect dysregulated emotional reactions to stress or provocation, a propensity to experience anger and frustration may be more strongly related to reactive, as compared to proactive, functions of relational aggression (see Frick & Morris, 2004). In fact, in one study with adults, hostility and anger were more strongly associated with reactive as compared to proactive relational aggression (MurrayClose et al., 2010). Additionally, Marsee and Frick (2007) reported that anger to provocation was uniquely associated with reactive, but not proactive, functions of relational aggression in a sample of adjudicated female adolescents (although see Ostrov et al., 2013). In contrast, a failure to experience fear, empathy, and guilt may lower inhibitions against the use of aggression for instrumental purposes (i.e., proactive aggression). Consistent with this suggestion, Marsee and Frick (2007) found that CU traits were more strongly related to proactive relational aggression than reactive relational aggression. Some researchers have also provided evidence that emotional processes are more strongly associated with relational forms of aggression in females than in males; for instance, Marsee, Silverthorn, and Frick (2005) found that CU traits were positively correlated with relational aggression in female, but not male, fifth to ninth graders. However, findings regarding gender differences have been mixed; in fact, some studies have failed to find evidence of gender differences (Czar et al., 2011), and others have reported that emotional processes are more strongly associated with relational aggression in males than in females (e.g. Loudin, Loukas, & Robinson, 2003). Additional research is necessary to explore the conditions under which distinct emotional processes translate into relationally aggressive conduct.

Risk Factors: Social Processes Parenting and Attachment Developmental psychopathologists have considered multiple social risk factors in the development of relationally aggressive conduct, including familial processes such as parenting and attachment. In a recent meta-analytic review of research in this area, Kawabata, Alink,

Tseng, van IJzendoorn, and Crick (2011) identified four clusters of parenting that may play an important role in the development of relational forms of aggression: psychological control, negative/harsh (coercive), uninvolved (permissive), and positive parenting. A comprehensive discussion of the parenting factors associated with relational aggression is beyond the scope of this chapter and available elsewhere (e.g., Kawabata et al., 2011); however, we review some of the main theoretical and empirical findings in this area. Importantly, many of the studies described next have considered both mother and father influences as well as the potential role of child gender. Although the vast majority of studies concern the relational aggression construct, a handful of studies feature indirect or social aggression. Psychological Control Many of the early studies of the role of parenting behaviors in the development of relational aggression are based on the supposition of social learning theory that parents influence their children's social behaviors via modeling (Kawabata et al., 2011). From this theoretical perspective, studies explored whether relational aggression in children might mirror parental strategies like psychological control (e.g., threats of love withdrawal; see Nelson & Crick, 2002, for details). In one of the first studies in this area, Hart, Nelson, Robinson, Olsen, and McNeilly-Choque (1998) reported that relational aggression was not associated with a small set of psychological control items in a sample of Russian preschoolers. However, when the full complement of psychological control items in this data set were considered dimensionally, a number of important associations emerged (Nelson, Yang, Coyne, Olsen, & Hart, 2013). Casas et al. (2006) also documented the utility of a dimensional approach to psychological control in predicting early childhood relational aggression with a U.S. sample. In an interesting extension of this work, Nelson, Hart, Yang, Olsen, and Jin (2006) explored how combined and differential parenting by mothers and fathers predicted relational aggression in a sample of Chinese preschoolers. A latent sum and difference structural equation model showed that combined psychological control (i.e., cumulative engagement in psychological control by mothers and fathers) was associated with heightened levels of relational aggression in girls. Interestingly, the extent to which fathers engaged in higher levels of psychological control was also positively associated with relational aggression in girls. In contrast, psychological control was not related to relational aggression in boys. These gender differences mirror findings reported by Nelson and Crick (2002) indicating that paternal psychological control predicted relational aggression in daughters in a sample of third graders. Associations between psychological control and relational aggression have emerged in studies across developmental periods. For instance, Gaertner and colleagues (2010) reported that psychological control predicted relational aggression among a group of children ages 9 to 12 years. In a Belgian sample of 8- to 10-year-old children, Kuppens, Grietens, Onghena, and Michiels (2009a) found support for the specialized associations between parental control and child relational aggression proposed by Nelson and Crick (2002). Specifically, parental psychological control was associated with child relational, but not physical, aggression. Soenens, Vansteenkiste, Goossens, Duriez, and Niemiec (2008) also demonstrated with a Belgian adolescent sample that parental psychological control (by fathers or mothers)

predicted increased adolescent relational aggression. Some evidence indicates that parental psychological control predicts relationally aggressive behavior across relational contexts (e.g., in dating relationships, Leadbeater, Banister, Ellis, & Yeung, 2008; in sibling relationships, Campione-Barr, Lindell, Greer, & Rose, 2014). Kawabata and colleagues (2011) provided meta-analytic evidence that paternal, but not maternal, psychological control was positively associated with relational aggression. In their more recent meta-analysis, Kuppens, Laurent, Heyvaert, and Onghena (2013) reported a positive yet weak association between parental psychological control and child relational aggression. Taken together, findings are consistent with the hypothesis that parental psychological control contributes to child relational aggression, although some evidence suggests that effects more often emerge in the context of paternal, rather than maternal, behaviors. Negative/Harsh (Coercive) Parenting According to social learning theory, negative or harsh parenting practices may teach children that aversive and aggressive behaviors, including relational aggression, are an effective strategy to get one's own way (Nelson & Crick, 2002). From this perspective, modeling of negative parent behaviors may lead children to engage in novel aggressive behavioral patterns (e.g., parent hostility may incite child relational aggression; Kawabata et al., 2011). Negative parenting behavior may also interfere with the establishment of secure attachment relationships, ultimately increasing risk for involvement in relational aggression. Researchers have explored a variety of types of negative parenting behaviors, including parent-child conflict, coercion, inconsistent discipline, parental negativity, and child maltreatment (see Kawabata et al., 2011). Findings from several studies suggest that negative parenting is associated with children's engagement in relational aggression during early childhood. In early research testing these hypotheses, Hart et al. (1998) reported that maternal physical/verbal coercion was positively associated with relational aggression in a sample of Russian preschool children. More recently, Ostrov and Bishop (2008) assessed whether general indicators of parent-child conflict were associated with teacher ratings or observations of preschool aggression. Results showed that, for both teacher ratings and observations, conflict uniquely predicted relational aggression above and beyond physical aggression and gender. Although findings are mixed (A. Russell, Hart, Robinson, & Olsen, 2003; Sandstrom, 2007), some researchers have also provided evidence indicating that authoritarian parenting is associated with relational aggression (e.g., Casas et al., 2006). Similarly, McNamara, Selig, and Hawley (2010) documented that mothers who provided little warmth and autonomy support, in combination with highly restrictive control, had young children who were more (relationally) aggressive. Similar findings have emerged in older children. For instance, Nelson and Crick (2002) found that maternal coercive control predicted relational aggression in third-grade children. Brown, Arnold, Dobbs, and Doctoroff (2007) videotaped mother-child interactions and coded for parental overreactivity (harsh/coercive) and negative maternal affect in a sample 5- to 8-yearold children who were predominantly European American or Puerto Rican. Among the European American participants, negative affect and overreactivity predicted heightened

levels of relational aggression. In the multivariate context, negative affect proved more important than other parenting variables. However, no significant associations emerged among the Puerto Rican dyads (n = 18). Replication with larger samples is necessary to determine whether Puerto Rican children react differently to parenting than European American children. Although limited research has examined the role of child maltreatment in the development of relational aggression, preliminary findings suggest that more severe instances of negative parenting, such as abuse, may portend risk for relational aggression. Specifically, CullertonSen et al. (2008) found that a documented history of child maltreatment was associated with elevated levels of relational aggression among youth ages 6 to 12 years. Interestingly, child maltreatment was associated with relational aggression for girls only, and findings primarily emerged in the context of sexual abuse. In a study with adults, a history of abuse was associated with heightened levels of reactive and romantic relational aggression (MurrayClose et al., 2010). These results highlight the potential role of severe forms of negative parenting, such as abuse, in the development of relational aggression. Several studies have also examined the association between early negative parenting and trajectories of indirect or social aggression. For example, Browne, Odueyungbo, Thabane, Byrne, and Smart (2010) found that hostile-ineffective parenting, wherein parents demonstrate annoyance, antagonism, and mood-dependent behavior, increased the risk of a child's development of indirect aggression over the course of 6 years. Vaillancourt, Miller, Fagbemi, Côté, and Tremblay (2007) examined predictors of different developmental trajectories of indirect aggression in children from age 2 to age 10. They identified two groups: increasing users (35%) and stable low users (65%). For boys only, parental inconsistency (lack of follow-through) at age 2 predicted an increased likelihood of belonging to the increasing group. Using the same data set, Côté et al. (2007) assessed the joint development of physical and indirect aggression from age 2 to age 8. Results showed that hostile parenting at age 2 predicted high levels of both physical and indirect aggression over time. Underwood et al. (2009) examined six different developmental trajectories for social and physical aggression, following children from age 9 to age 13. Maternal authoritarian parenting predicted children's membership in the high increaser group. Meta-analytic findings further underscore the important role of negative parenting in the development of relational aggression: Kawabata and colleagues (2011) reported that both maternal and paternal negative parenting were positively associated with relational aggression. Uninvolved (Permissive) Parenting Theorists have also hypothesized that permissive parenting may be associated with elevated levels of child relational aggression; from this perspective, parents may fail to intervene when their children are aggressive, thereby reinforcing these undesirable behaviors (Casas et al., 2006). Consistent with this hypothesis, Casas et al. (2006) found that permissive parenting styles were associated with child relational aggression. Similarly, Leadbeater et al. (2008) reported that low levels of parental monitoring were associated with greater use of relational aggression in the context of dating relationships among 12- to 18-year-old adolescents. In Underwood and colleagues' (2009) examination of joint developmental trajectories of social

and physical aggression across age 9 to age 13, maternal permissive parenting predicted children's membership in the high increaser and medium increaser groups. Some evidence suggests that permissive parenting behaviors may be more strongly associated with relational aggression in girls as compared to boys. For instance, Sandstrom (2007) reported that maternal permissiveness was positively associated with relational aggression in 9- to 11-year-old girls but was unrelated to such conduct in boys. In observational work in this area, Brown et al. (2007) documented that videotaped mother-child interactions characterized by laxness were positively associated with relational aggression for girls but negatively associated for boys. However, meta-analytic findings do not suggest gender moderation; instead, findings indicate that both maternal and paternal permissiveness are positively associated with relational aggression in both boys and girls (Kawabata et al., 2011). Positive Parenting Researchers have hypothesized that positive parenting, including secure attachment relationships, authoritative parenting styles, and positive affect, may reduce children's risk for involvement in relational aggression (Kawabata et al., 2011). In early international research in this area, Hart et al. (1998) reported that paternal responsiveness (an authoritative dimension) was most predictive of Russian preschoolers' relational aggression, even when controlling for other parenting dimensions (although see Casas et al., 2006). Casas et al. (2006) were the first to assess whether attachment security was related to relational aggression in a preschool sample (see Michiels, Grietens, Onghena, & Kuppens, 2008, for a theoretical synopsis regarding attachment). They found the association to be specific to gendered dyads. Whereas relationally aggressive girls were less likely to have a secure attachment with their mother, relationally aggressive boys were less likely to have a secure attachment with their father. Other positive parenting behaviors have also been shown to predict reduced risk for involvement in relationally aggressive conduct. For instance, Brown et al. (2007) reported that more positive affect observed in mother-child dyadic interactions was associated with less relational aggression. In addition, parent-child bonding has been found to predict lower levels of social aggression (Karriker-Jaffe et al., 2013). Browne et al. (2010) reported that parental consistency (parental follow-through with discipline and requests) and positive interaction (parental praise/amount of quality time with child) acted as buffers against the development of indirect aggression over the course of 6 years. Similarly, Vaillancourt et al. (2007) found that more positive parent-child interactions at age 2 reduced the likelihood that boys would exhibit increasing trajectories of indirect aggression from age 2 to age 10. Finally, parental control, including consistent rules and monitoring, have been associated with lower levels of social aggression (Karriker-Jaffe, Foshee, Ennett, & Suchindran, 2013). Meta-analytic findings have confirmed the important role of positive maternal and paternal parenting behaviors in child involvement in relational aggression (Kawabata et al., 2011). Directions for Future Research Despite the burgeoning literature regarding parenting correlates of relational aggression, several critical questions remain. First, it will be important for researchers to examine the

mechanisms that link parenting practices to child relational aggression; for instance, aversive parenting may enhance a child's HAB (Nelson & Coyne, 2009), which may in turn be linked to relational aggression. In addition, positive parenting may promote adaptive child behaviors, such as disclosure (e.g., not keeping secrets from one's parents), that prevent relationally aggressive conduct. In fact, higher levels of relational aggression have been reported among youth who engage in lower levels of disclosure (Gaertner et al., 2010) and higher levels of concealment (Leavitt, Nelson, Coyne, & Hart, 2013) in their interactions with their parents. Moreover, child disclosure may serve as a stronger predictor of adolescent antisocial conduct than parental monitoring (Criss et al., 2015). It may also be useful for researchers to examine interactions among parenting behaviors; in fact, in one study, the combination of high parental solicitation and psychological control predicted increased engagement in child relational aggression (Gaertner et al., 2010). Moreover, the majority of studies of parenting and relational aggression either make use of concurrent data or, if they are longitudinal, do not attempt to ascertain reciprocal effects between parenting and child aggression. This limitation is significant, given evidence of evocative gene-environment correlations in behaviors such as aggression (Brendgen, 2012). In one of the only studies to assess bidirectional associations over time, Kuppens, Grietens, Onghena, and Michiels (2009c) used a cross-lagged panel design (three measurement points at 1-year intervals) with a middle-childhood sample. Parents and children reported on parental psychological control, and child aggression was reported by parents, peers, and teachers. Reciprocal effects emerged for mother-child dyads, whereas father-child dyads were best characterized by unidirectional parent effects, at least when parents reported on both psychological control as well as aggression. Future research may also benefit from greater attention to cultural differences in associations between parenting behaviors and relational aggression. Some evidence suggests that cultural factors may influence the degree to which parents adopt psychologically controlling behavior. Using longitudinal data collected in seventh grade and ninth grade, Shuster, Li, and Shi (2012) found that, controlling for Time 1 physical and relational aggression, maternal endorsement of the cultural value of social harmony predicted less engagement in parental psychological control, which in turn was associated with Time 2 relational aggression. These findings highlight the potential role of cultural values in parenting practices and, ultimately, children's engagement in relational forms of aggression.

Parental Beliefs and Expectations Although most research regarding the association between parenting and child relational aggression has focused on parenting behaviors, a few studies have also highlighted the important role of parent expectations in children's aggressive behavior. Murray, Haynie, Howard, Cheng, and Simons-Morton (2010) conducted a unique study with African American sixth graders living in an urban community plagued with violence. Over the course of 3 months, they assessed how initial parental expectations and practices predicted overt and relational aggression at school. Multivariate analyses demonstrated that parental expectations that their child would engage in peaceful solutions to conflict (a) buffered against overt aggression when parental support and responsiveness was relatively low and (b) buffered against both overt

and relational aggression in the context of high psychological control by parents. Several studies have examined maternal beliefs about relational aggression and how these beliefs translate into discipline responses, child beliefs, and child behavior. For instance, in a study of parents of preschoolers, Werner, Senich, and Przepyszny (2006) found that, relative to physical aggression, mothers reported they would be less upset, angry, and sad in response to their child's engagement in relational aggression. Accordingly, relational aggression was also perceived as less deserving of parental intervention. However, mothers who said that they would intervene by helping their child understand that he or she had violated a social or moral convention had children who were less relationally aggressive. Similarly, Werner and Grant (2009) found that mothers perceived relational aggression as more normative and acceptable than physical aggression and their children to be less responsible for their engagement in such behavior. Children's perceptions of relational aggression tended to correspond with those of their mothers. The more mothers considered relational aggression to be problematic, however, the more they reported they would use sternness and disapproval in response to their daughters' relational aggression. Daughters of more stern mothers were in turn more likely to be prosocial and better accepted by peers (these associations did not emerge for boys). Although preliminary, findings from these studies highlight the critical need for research regarding the role of parental views of aggression in the development of child relational aggression.

Interparental Conflict Prior research has indicated that marital conflict is related to childhood adjustment difficulties, including physical aggression, particularly when child exposure to conflict is high (Cummings, 1994). A number of potential mechanisms could explain this association. From a social learning perspective, children may model interparental aggression, believing it is acceptable behavior. Parents may inappropriately interpret the intentions of one another (e.g., demonstrate HAB), and the child may adopt similar thinking in interpreting his or her own social experiences and thereby act more aggressively. Depending on each child's individual coping style, interparental conflict may also generate background anger, a stressor that heightens arousal in children and potentially elicits aggression in some children (Cummings, 1987). Finally, marital conflict may undermine parenting and the parent-child relationship, thereby leading to more child aggression. Few studies have assessed whether marital conflict is related to the development of relational aggression. In one notable exception, Hart and colleagues (1998) assessed overt marital conflict (e.g., verbal hostility, quarreling, and physical abuse) and marital exclusion (e.g., avoiding, ignoring, and withdrawing affection) in their Russian preschooler study. Results showed that overt marital conflict, but not marital exclusion, was positively associated with relational aggression among boys but not among girls. Other work, however, has documented effects for girls. Specifically, Underwood, Beron, Gentsch, Galperin, and Risser (2008) examined interparental conflict strategies (stonewalling, triangulation, verbal aggression, and physical aggression) in the prediction of children's social aggression. Notably, the interparental conflict strategies were collectively modeled in structural equation modeling as one latent construct (negative conflict strategies). Results showed that only the maternal

negative conflict strategies predicted social aggression, and findings emerged for daughters but not for sons. More recently, Karriker-Jaffe et al. (2013) reported that family conflict was positively associated with social aggression for both boys and girls, although analysis of trajectories of social aggression indicated that compared to boys from high-conflict families, girls from high-conflict families exhibited higher initial levels of social aggression (at age 11) and a more gradual decline in these behaviors across adolescence. Although preliminary, these findings highlight the potential role of marital relationships and interactions in child relational aggression. Future research in this area may be enhanced by more careful definitions of interparental relational aggression, such as the distinction between direct and indirect forms of relational aggression. Consistent with this idea, Carroll and colleagues (2010) developed a measure reflecting two subtypes of marital relational aggression: social sabotage (indirect) and love withdrawal (direct). Study results showed that wives were more likely to be relationally aggressive than husbands, and such tactics were concurrently associated with lower marital quality and higher marital instability. The next step is to determine whether these forms of marital relational aggression are uniquely predictive of relational aggression in children. Given that children are not likely to witness social sabotage or some forms of love withdrawal, the negative impact may occur via mechanisms other than direct observation of these subtle behaviors, such as impaired parenting or witnessing cooccurring overt marital conflict. Future studies must attend to the specific processes that link interparental conflict and child relational aggression, such as parenting behaviors. For instance, Li, Putallaz, and Su (2011) have assessed how interparental conflict styles may indirectly predict child overt and relational aggression via coercive or psychological control. Using a sample in Beijing, China, they found that fathers' overt interparental conflict was associated with boys' aggression, and this relation was mediated by paternal coercive control. Maternal covert conflict was positively related to boys' aggression, and this association was mediated by maternal psychological control. Given evidence of differences in effects based on both parent and child gender (e.g., Li et al., 2011), future work in this area should also carefully evaluate the role of gender in the processes linking interparental conflict and child relational aggression.

Sibling Relationships By middle childhood, time spent with siblings tends to outweigh other relationships (McHale & Crouter, 1996); thus, there is good reason to suppose that siblings may substantially affect each other's involvement in relationally aggressive conduct. Sibling relationships differ significantly from peer acquaintances or friendships in ways that may promote the development of relational aggression. For instance, the greater intimacy of family relationships allows siblings access to a great deal of private and potentially damaging information about each other (precisely the type of information that would lend itself to relational aggression). Additionally, as sibling relationships are involuntary, siblings must learn to deal with each other's eccentricities, potentially leading to relatively high levels of conflict. Sibling interactions may also be influenced by the nature of marital or parent-child relationships, reflecting the systemic workings of a family. As briefly noted earlier, evidence indicates that preschoolers engage in

more relational aggression with siblings than with friends. Following these children and their siblings from age 4 to age 8, Stauffacher and DeHart (2006) observed that by the time the target child was 8 years old, sibling and friend dyads no longer differed in their levels of relational aggression. Specifically, over the 4-year period, friend dyads increased in their relational aggression, whereas the same behavior between siblings declined. Several factors, including child gender and age, may influence the frequency and type of relational aggression used against siblings. With respect to frequency, Campione-Barr et al. (2014) found that, within the context of sibling dyads, younger siblings were more often the target of sibling relational aggression than older siblings. With respect to type of relational aggression, Stauffacher and DeHart (2005) found that girls with younger siblings tended to use nonverbal (ignoring) strategies with siblings, whereas girls with older siblings tended to focus on more verbal strategies. Boys, in contrast, preferred verbal forms of relational aggression, regardless of the age of the sibling. Understandably, sibling relational aggression is tied to other aspects of the sibling relationship. In particular, relational aggression is more likely if there is less emotional support and more negativity between siblings, regardless of the gendered nature of the dyad (same versus opposite sex) or the age of siblings (Updegraff, Thayer, Whiteman, Denning, & McHale, 2005). Ostrov, Crick, and Stauffacher (2006) suggested that use of relational aggression in the context of sibling relationships may provide a training ground for younger siblings to learn aggressive behavioral strategies; from this perspective, youth may model their older siblings' relational aggression and thus may be at increased risk for engaging in these behaviors in the peer context. These authors observed both siblings of a given dyad in the same preschool environment (the younger sibling enrolled 1 to 2 years later). Observations showed that older sisters were more relationally aggressive than older brothers. Older siblings were also more aggressive than younger siblings with same-sex peers. Consistent with expectations, older siblings' relational aggression toward peers was predictive of the same behavior in younger siblings (this was also found for peer-directed physical aggression). Only two studies have considered how parent-child dynamics may be connected to sibling relational aggression (Campione-Barr et al., 2014; Updegraff et al., 2005). In the first, Updegraff and colleagues (2005) reported that when parents provided less emotional support to their children, sibling relational aggression increased. Less time with fathers also predicted higher levels of sibling relational aggression. Interestingly, parents who intervened in sibling conflict with authoritarian practices (e.g., punishments) tended to have children, particularly daughters, who engaged in relational aggression. In the second study, Campione-Barr et al. (2014) found that maternal psychological control predicted heightened use of sibling relational aggression in a sample of adolescents. Although the findings from these studies may suggest that inadequate parenting contributes to sibling relational aggression, another possibility is that parents are reacting to aggressive siblings with less warmth and involvement and greater psychological control. Finally, differential treatment of siblings by parents is relevant: Children report less sibling relational aggression when parents are perceived to be evenhanded (Updegraff et al., 2005). However, work in this area is still in its infancy, and additional research is needed to identify the developmental precursors and outcomes

associated with use of relational aggression in the sibling context.

Peer Relationships Much of the extant relational aggression research has focused on these behaviors in the context of the peer group, and several peer processes have been identified that may serve as risk factors in the development of relational aggression. More specifically, researchers have suggested that peer status, peer victimization, and quality of close dyadic relationships may all contribute to the emergence of relational forms of aggression. Peer Status One of the most robust correlates of relational aggression across developmental periods is low peer status, including low preference (e.g., Crick & Grotpeter, 1995), and its associated components of high rejection (i.e., frequently being identified by peers as disliked) and low acceptance (i.e., rarely being identified by peers as liked). Most researchers in this area have proposed that engagement in relational aggression leads to poor peer status (see the “Developmental Outcomes: Maladaptive Correlates” section); however, the vast majority of this work has been cross-sectional in nature, and there is reason to expect that low status in the peer group may also precede the use of relationally aggressive behaviors. One influential causal model of the relation between low preference and aggression suggests that low preference leads to a constrained pool of potential partners comprised primarily of peers who also have impaired social skills (Bierman, 2004). This smaller peer group likely contributes to the behavioral homophily (associating with other similar peers) often observed among aggressive children (Parker, Rubin, Price, & DeRosier, 1995). Perhaps more important, this process contributes to a negative spiral whereby rejected children become more and more socially incompetent relative to their well-accepted peers, promoting higher levels of both physical and relational aggression over time (Bierman, 2004). Recent research has documented both homophily and socialization of relational aggression in peer groups (Sijtsema, Ojanen, et al., 2010), suggesting that this may be a plausible mechanism involved in the development of relational aggression. In fact, although evidence is mixed (e.g., Orue & Calvete, 2011; Ostrov et al., 2013; Tseng, Banny, Kawabata, Crick, & Gau, 2013), several researchers have found that indices of low preference are longitudinally associated with future relational aggression. For instance, Zimmer-Gembeck, Crick, and Geiger (2005) reported that low social preference predicted increases in relational aggression across 3 years, although in this study, effects were significant only for boys. In addition, Werner and Crick (2004) reported that rejection by peers predicted elevated levels of relational aggression one year later in a sample of elementary school children. Murray-Close and Crick (2006) also found that increases in peer rejection over time were dynamically associated with increases in relationally aggressive behaviors. However, other researchers have suggested that high peer status may serve as a social risk factor for the development of relationally aggressive conduct. Foundational to much of this research is the distinction between perceived popularity (based on most and least popular

nominations by peers) and preference. Perceived popularity is an indicator of children's social status and impact whereas preference is an indicator of children's likability (Parkhurst & Hopmeyer, 1998). Perceived popular youth benefit from possessing what is described as “the expressive equipment of popularity”: physical attractiveness, athleticism, and evidence of significant material wealth (Lease, Kennedy, & Axelrod, 2002, p. 87; see also Rosen & Underwood, 2010). In short, these youth have superficial traits that make them well positioned for popularity. Some researchers have argued that perceived popular children employ relationally aggressive behavior in efforts to maintain their high social status and power (Sandstrom & Cillessen, 2006). Consistent with this hypothesis, findings from several studies indicate that high levels of perceived popularity are related to increases in relational aggression over time (Rose, Swenson, & Waller, 2004; Tseng et al., 2013), particularly among girls (Cillessen & Mayeux, 2004). Interestingly, some evidence suggests that a curvilinear effect best represents the association between popularity and relational aggression, such that elevated engagement in relational aggression is evident in youth with both high and low levels of popularity (Prinstein & Cillessen, 2003). Moreover, although the majority of research suggests that indices of low preference serve as a risk factor for relational aggression, findings are mixed. For instance, some work indicates that high social acceptance is related to increases in relational aggression over time among youth (Tseng et al., 2013), especially girls (Orue & Calvete, 2011; see also Kuppens, Grietens, et al., 2009b). However, it is possible that positive associations between preference and the development of relational aggression are an artifact of the overlap between preference and popularity. Consistent with this notion, Prinstein and Cillessen (2003) found that when both popularity and preference were included in one model, social preference was related to decreases in relational aggression over time among girls (see Adams, Bartlett, & Bukowski, 2010, for evidence that social dominance may serve as a suppressor variable in the association between low preference and heightened relational aggression). These findings suggest that relationally aggressive behaviors may be more strongly linked to high peer status when measures focus on perceived popularity as compared to preference. These results also raise the possibility that low and high peer status may promote relational aggression for different reasons; specifically, low preference or unpopular youth may use relational aggression to deal with negative experiences in the peer group, whereas popular children may engage in relational aggression because their high social status provides them with social power for the effective use of these behaviors (see Kuppens, Grietens, et al., 2009b) and because peers fail to censure them when they engage in such conduct (Rose, Swenson, & Waller, 2004). These findings also suggest that researchers interested in investigating the role of peer status in the development of relational aggression should include multiple facets of status, including preference and popularity, so that the unique role of each can be assessed (e.g., Bowker, Ostrov, & Raja, 2012). There are several important directions for future research regarding the processes underlying the association between peer status and the development of relational aggression. First, it will be important for researchers to assess the extent to which relations between peer status and relational aggression are incidental such that they reflect the action of third variables (see

Rubin, Bukowski, & Parker, 2006). For example, depressive symptoms may function as a third variable that promotes both rejection by peers as well as aggressive behavior (e.g., Agoston & Rudolph, 2013). As another example, relational aggression may be associated with poor functioning in the peer group as a result of its overlap with other problematic behaviors, such as physical aggression. In fact, in one study, the relation between relational aggression and lower acceptance became nonsignificant when overt aggression was controlled (Smith, Rose, & Schwartz-Mette, 2009). However, some research indicates that peer rejection longitudinally predicts relational aggression, even when controlling for physical aggression (Murray-Close & Crick, 2006), suggesting that negative peer relationships may be uniquely associated with relational aggression. It will also be important to examine potential third variables in the relation between high peer status and relational aggression; for instance, positive child characteristics, such as social skills, may promote both positive peer outcomes (e.g., acceptance, popularity) as well as relational aggression (Orue & Calvete, 2011). Second, future research should test transactional models (Sameroff & MacKenzie, 2003) of the relations between social status and relational aggression. Transactional views of development suggest bidirectional and mutually influencing correlations between key variables that in combination may best explain a developmental pathway. From this perspective, peer status may interact with child characteristics to promote involvement in relational aggression. For example, it is possible that peer rejection in conjunction with maladaptive social informationprocessing patterns (e.g., HAB; see the “Risk Factors: Cognitive and Emotional Processes” section) better predicts relational aggression than either of these factors individually (see the “Moderation Models” section). Given evidence of gender differences in the association between peer status and later relational aggression (e.g., Cillessen & Mayeux, 2004), it will also be important for future research to evaluate the extent to which peer status differentially predicts relational aggression in boys versus girls. Recent cognitive models of relational aggression (see the “Toward an Integrated Gender-Linked Model of Aggression” section) highlight the important role of gender schemas in children's aggression; from this perspective, peer status may most strongly predict relational aggression in girls because youth select aggressive behavior patterns that are consistent with their gender schemas. Several additional moderators of the association between relational aggression and popularity merit empirical attention. For instance, a relatively egalitarian classroom status hierarchy may promote relationally aggressive strategies among higher-status youth because it increases their need to defend their desirable social positions (Zwaan, Dijkstra, & Veenstra, 2013). Additionally, Mayeux and Cillessen (2008) found that perceived popular adolescents who also knew they were popular exhibited the highest concurrent levels of peer-reported aggression (relational aggression for girls and overt aggression for boys). Awareness of high status may provide license for aggressive behavior or alternatively attune the adolescent to the need to fight to maintain status. Finally, given the theoretical processes proposed to underlie relations between peer status and later relationally aggressive conduct, researchers should examine whether associations depend on whether the aggression is proactive versus reactive in function. Insofar as rejected youth use relational aggression to cope with negative peer experiences, they may be more likely to engage in reactive relational aggression. In contrast, if

popular youth use relational aggression to maintain their high social status, they may be particularly likely to engage in proactive relational aggression. Consistent with this suggestion, in one study, proactive relational aggression was positively associated with popularity but not preference (Prinstein & Cillessen, 2003). However, reactive relational aggression was positively related to both popularity and preference (Prinstein & Cillessen, 2003), perhaps because high-status youth used relational aggression to defend challenges to their coveted social positions. Longitudinal research is necessary to further explore how peer status relates to the development of proactive and reactive functions of relational aggression. Victimization Relational aggression and victimization have been found to be moderately correlated at all points in development (e.g., Crick et al., 2001), and evidence indicates that a subset of relational perpetrators are also victims (Gros, Gros, & Simms, 2010). Engaging in relational aggression may lead to tit-for-tat responses whereby the relational aggressor may also become a victim (see the “Developmental Outcomes: Maladaptive Correlates” section). However, it is also plausible that victimization serves to elicit relationally aggressive responses over time (e.g., Ostrov, 2010; Ostrov & Godleski, 2013). In fact, when asked how they respond to experiences of victimization, middle school and high school students commonly identified aggressive strategies (Waasdorp & Bradshaw, 2011). Researchers have proposed several mechanisms whereby experiences of peer victimization may promote aggressive behavior. Victimized youth may engage in aggression in an effort to retaliate against their tormentor or to prevent future victimization experiences (Yeung & Leadbeater, 2007). The experience of being a frequent target of relational victimization may result in feelings of isolation, loneliness, and low self-esteem (Gomes, Davis, Baker, & Servonsky, 2009). These adjustment difficulties may lead relational victims to lash out at their tormentors in a fashion similar to what they have experienced. Some evidence also suggests that experiences of relational victimization are associated with heightened HAB for relational provocations (Hoglund & Leadbeater, 2007; Yeung & Leadbeater, 2007), which in turn has been found to predict involvement in relational aggression in some studies (see the “Risk Factors: Cognitive and Emotional Processes” section). Additionally, similar to their rejected peers, victimized youth may have a limited pool of social interaction partners, thus reducing opportunities for learning positive social behaviors (Hanish & Guerra, 2002). Several researchers have provided evidence consistent with the suggestion that peer relational victimization predicts increased relationally aggressive conduct. Sullivan, Farrell, and Kliewer (2006) reported that relational victimization was associated with heightened levels of concurrent relational aggression in a sample of predominantly African-American eighth graders; interestingly, although effects emerged for both boys and girls, associations were stronger for boys. Longitudinal work has provided evidence that experiences of relational victimization may promote relationally aggressive conduct over time. For instance, Yeung and Leadbeater (2007) reported that relational victimization was related to preadolescents' engagement in relational aggression 5 months later. Sugimura and Rudolph (2012) found that relational victimization in second grade was associated with engagement in relational

aggression in third grade. Finally, using data from the NICHD SECCYD, Ostrov and Godleski (2013) reported that relational victimization in third grade was associated with relational aggression in sixth grade. An important question is whether the experience of any type of peer maltreatment predicts increased levels of relationally aggressive conduct or whether there is specificity between the type of victimization and children's behavioral responses. Theoretically, there is reason to expect that experiences of relational victimization will be particularly likely to elicit relationally aggressive responses. For instance, Yeung and Leadbeter (2007) suggested that relational victimization may be more likely than other forms of victimization to promote HAB for relational provocations and thus relationally aggressive behaviors. It is also possible that youth learn to engage in aggressive behaviors by modeling their tormentors' conduct (Ostrov, 2010). Consistent with these suggestions, Ostrov (2010) reported that teacher-reported relational victimization predicted increases in observed relational, but not physical, aggression across the school year in a sample of preschoolers. In contrast, physical victimization predicted increases in physical aggression in this study. However, several of the theoretical processes proposed to underlie the association between victimization and aggression (e.g., using aggression to prevent future victimization; Yeung & Leadbeater, 2007) suggest that youth may use relational aggression in response to a variety of peer victimization experiences; thus, it will be important for future research to evaluate the relative influence of distinct subtypes of peer maltreatment in the development of relationally aggressive conduct. Preliminary research has tested some of the theoretical mechanisms proposed to account for the association between relational victimization and relationally aggressive conduct. For instance, Ostrov et al. (2011) provided evidence supporting a social processes model in which the desire for exclusivity with close friends and HAB for relational provocations mediated the association between relational victimization and relational aggression in a sample of female emerging adults. Similarly, Yeung and Leadbeater (2007) found that HAB for relational provocations partially mediated the association between relational victimization and concurrent relational aggression in a sample of preadolescents. However, significant mediation did not emerge in longitudinal analyses. Some researchers have further suggested that victimization may serve as a mechanism linking peer rejection and relational aggression. For instance, Adams et al. (2010) found that victimization by peers (a composite including both physical and relational forms of victimization) mediated the relation between low peer liking and relational aggression in fifth- and sixth-grade males and females. Moderation effects were also documented, such that peer dislike predicted relational aggression only among youth who were highly victimized. There are several directions for future work on the relation between peer victimization and the development of relationally aggressive conduct. First, researchers may benefit from evaluating the role of chronicity and severity of victimization in the development of relational aggression. In fact, preliminary evidence indicates that chronic victimization may be particularly likely to elicit relationally aggressive behaviors. Specifically, Rudolph, Troop-Gordon, Hessel, and Schmidt (2011) found that levels of victimization (a composite of both physical and relational victimization) in second grade and increases in these experiences into fifth grade predicted

heightened relational aggression in fifth grade (controlling for second-grade relational aggression). Interestingly, this association was stronger among girls than boys. These gender differences support the suggestion that girls who are targeted by negative treatment by peers may be particularly likely to select relationally aggressive responses because such conduct is consistent with their gender schemas (see the “Toward an Integrated Gender-Linked Model of Aggression” section). In regard to severity, it is possible that even infrequent episodes of relational victimization may elicit relationally aggressive responses if the victimization is very severe in nature. Moreover, given evidence that aggression and victimization by parents and siblings may also influence children's behaviors within the peer group (Casas et al., 2006; Stauffacher & DeHart, 2005), future research should consider the role of victimization across multiple contexts. Relational Aggression in Friendship Groups and Dyadic Contexts To date, the majority of research examining peer risk factors for the development of relational aggression has focused on processes occurring in the larger peer context (e.g., peer status; peer victimization). However, as relational aggression involves the manipulation of and damage to interpersonal relationships, it is likely that such conduct often emerges in the context of close, dyadic relationships (Grotpeter & Crick, 1996) and friendship groups (e.g., Espelage, Holt, & Henkel, 2003; Low, Polanin, & Espelage, 2013). In fact, relational aggression occurs in romantic relationships (Casas & Albers, 2014; Linder et al., 2002) and close friendships (Grotpeter & Crick, 1996) with some frequency. Youth with mutual antipathy relationships (i.e., enemy relationships in which both members of the dyad report disliking the other peer) also engage in heightened levels of relational aggression, perhaps in part because they use these behaviors to harm their disliked peers (Murray-Close & Crick, 2006). Thus, it is essential for researchers to understand the friendship group and dyadic processes that may contribute to the emergence of relationally aggressive conduct. Several researchers have suggested that the qualities and characteristics of close friends, such as their aggressive behavior and social status, may play an important role in the development of children's and adolescents' relationally aggressive conduct. In fact, one important mechanism argued to underlie the development of aggressive behavior problems is socialization by close friends in the context of friendship networks (Espelage et al., 2003; Sijtsema, Ojanen, et al., 2010). More specifically, researchers have documented that aggressive youth choose friends who are also highly aggressive (i.e., selection), and, over time, these youth exert mutual social influence on each other's aggression (i.e., socialization; see Espelage et al., 2003). From this perspective, a salient social risk factor for the development of relational aggression may be time spent with relationally aggressive friends. Moreover, relationally aggressive youth may be particularly likely to experience this risk factor because they select friends who are similar to themselves. In fact, researchers have provided evidence of both selection and socialization processes in relationally aggressive youths' friendship groups. For instance, Sijtsema, Ojanen, and colleagues (2010) found that relationally aggressive adolescents were more likely to become friends with relationally aggressive peers across one calendar year. Several potential

mechanisms may underlie these selection effects. For instance, relationally aggressive individuals may be viewed by their peers as undesirable social partners, and, as a result, their opportunities for friendships may be constrained to other aggressive peers (Parker et al., 1995). However, at least some relationally aggressive individuals appear to be socially attractive (Hawley, Little, & Card, 2007), and children may actively choose friends who are similar to themselves, given the salient role of similarity in interpersonal attraction (Simon, Aikins, & Prinstein, 2008). Interestingly, popularity may be the primary driver of friendship formation and aggression similarity a by-product of the association with popular friends (Dijkstra, Berger, & Lindenberg, 2011). An emerging body of research also provides support for the notion that youth are influenced by their friends' engagement in relational aggression. In one study, nonaggressive third-grade girls who were best friends with relationally aggressive girls became significantly more relationally aggressive themselves a year later (Werner & Crick, 2004). More recently, Sijtsema, Ojanen, et al. (2010) used social network analysis to document that young adolescents adopted relationally aggressive behavior patterns from friends, and findings persisted even when controlling for friendship selection effects. Similarly, Low and colleagues (2013) reported that sixth and seventh graders who were members of social networks that engaged in high levels of relational aggression exhibited increases in their own perpetration of such behavior across the course of one calendar year. Moreover, some evidence indicates that socialization processes may be more powerful in the development of relational forms of aggression than in physical aggression or physical fighting (Espelage et al., 2003; Low et al., 2013; Sijtsema, Ojanen, et al., 2010), underscoring the important role of friend and social network socialization in these behaviors. Researchers have identified several moderators of socialization processes in the development of relational aggression. For instance, some evidence indicates that socialization of relational aggression in friendship groups is especially likely when the child is not powerful or socially central; specifically, Lansford and colleagues (2009) found that in girls' cliques, peripheral group members were more likely than more central group members to engage in levels of indirect aggression consonant with the indirect aggression of the group leader. Socialization processes may also be most pronounced for youth who are already aggressive; consistent with this suggestion, Simon et al. (2008) reported that only sixth to eighth graders who were highly relationally aggressive and in romantic relationships with relationally aggressive partners exhibited increased involvement in such conduct over time. Several studies have also provided evidence indicating that friends' social status or social centrality may be associated with youth's relationally aggressive conduct. For instance, Peters, Cillessen, Riksen-Walraven, and Haselager (2010) found that girls with popular best friends were more likely to engage in relationally aggressive behavior. The authors suggest that popular friends may form a coalition and engage in relational aggression against others in an effort to gain or maintain high status. However, other research has highlighted the role of friendships with low-status peers in the development of relational aggression. More specifically, Neal and Cappella (2012) have shown that relationally aggressive children tend to have high social network centrality (in terms of the number of relationships), but only if

these relationships are with poorly connected peers. Therefore, children's relational aggression is enhanced when they are connected to many peers who are dependent on them for social resources. These authors suggest that relational aggression may be particularly effective when used against friends who do not have alternative options for obtaining relational benefits such as social support and companionship. Taken together, findings raise the possibility that having low-status friends promotes relational aggression against these same friends, whereas having high-status friends promotes the use of relational aggression against other members of the peer group. Some evidence also suggests that relationship qualities may play a role in the development of relational aggression. Specifically, relative to nonaggressive youth, relationally aggressive children's friendships are characterized by high levels of intimacy (closeness and selfdisclosure), conflict, betrayal, relational aggression within the relationship, and exclusivity (Cillessen, Jiang, West, & Laszkowski, 2005; Crick, Murray-Close, Marks, & MohajeriNelson, 2009; Grotpeter & Crick, 1996). This particular cluster of relationship qualities is problematic because it suggests that children with relationally aggressive friends are divulging highly personal information and secrets, either elicited by the relationally aggressive friend or by choice, and that this information can be used against them when relationship problems arise. Although most work in this area has been cross-sectional, results from one longitudinal study indicate that increases in intimate disclosure by a friend are associated with increases in relational aggression for girls only (Murray-Close et al., 2007). These findings are provocative because they suggest that some friendship characteristics that are frequently viewed as positive (e.g., intimacy) may at times facilitate the use of relational aggression.

Media Exposure The association between violent media exposure consumption and physically aggressive behavior has been thoroughly reviewed in the literature (see Anderson & Bushman, 2001; Anderson et al., 2010). Numerous theoretical models help to explain how exposure to aggressive media models can increase the likelihood of aggressive behavior over time (e.g., general aggression model [GAM], Carnagey & Anderson, 2003; script theory, Huesmann, 1986). The theoretically predicted links between exposure to violent media or aggressive content and aggressive behavior in children/adolescents are robust across types of methodology (i.e., longitudinal and cross-sectional studies; experimental and correlational studies) and cultures (see Anderson & Bushman, 2002). These effects have also been successfully extended to the investigation of nonphysical forms of aggression (i.e., relational and indirect aggression) in several studies (e.g., Archer & Coyne, 2005; Coyne et al., 2008; Gentile, Mathieson, & Crick, 2011). For example, Gentile, Coyne, et al. (2011) recently documented that school-aged children's exposure to media violence early in the academic year was associated with higher relational aggression (as well as physical and verbal aggression and low levels of prosocial behavior) later in the school year. Media exposure is also related to use of relational aggression in the context of romantic relationships; in one study, Coyne et al. (2011) found that exposure to relational aggression in the media was associated with romantic relational aggression for both young adult men and women.

There is also interesting evidence for what is referred to as the crossover effect (see Coyne, Archer, & Eslea, 2004; Coyne et al., 2008) in which viewing one type of media violence (e.g., physical) can lead to other types of aggression (e.g., relational) among viewers. For example, in a well-known longitudinal study, Huesmann, Moise-Titus, Podolski, and Eron (2003) found that girls who viewed physical violence on television as children engaged in more indirect aggression as adults. Ostrov, Gentile, and Crick (2006) documented a similar process among young preschool-age children. The GAM (see Carnagey & Anderson, 2003) predicts that media violence does not just influence aggression via social learning and modeling; rather, the exposure increases aggressive affect, arousal, and cognitive scripts (Ostrov, Gentile, & Mullins, 2013). Scripts or cognitive knowledge structures provide heuristics for how to act in given situations (Huesmann, 1986). Long-term exposure to aggressive media models can lead to aggressive behavior scripts and schemas as well as desensitization to aggressive or violent episodes (Carnagey & Anderson, 2003; Smith & Donnerstein, 1998). Viewing large amounts of relationally aggressive media may be associated with beliefs about the acceptability of relational aggression. A recent study showed that school-age children who viewed relationally aggressive television and movie models exhibited greater subsequent approval of relational aggression (i.e., normative beliefs about the acceptability of relational aggression) if they also had parents who did not actively mediate or discuss and scaffold the programming during the exposure (Linder & Werner, 2012). Finally, one recent experimental study documented that exposure to television clips depicting relational aggression was associated with HAB for ambiguous relational (but not instrumental) provocation scenarios (Martins, 2013); thus, HAB may serve as a cognitive mechanism linking aggressive media exposure and relational aggression. Recent work has also demonstrated how exposure to other forms of media content, including educational programming, may serve as a risk factor for relational aggression. For example, Ostrov et al. (2013) showed that exposure to educational media was associated with increases in parent-reported, teacher-reported, and observed relational aggression in a 3- to 5-year-old sample. These effects persisted over more than a 2-year period (Ostrov et al., 2013). The authors believed that children are modeling the aggressive behavior that is often shown as part of a friendship problem in these social-emotional-based programs. That is, although these programs are designed to ultimately teach social skills to children, young children may fail to connect the lesson or moral that is contained at the end of the 15- or 30-minute program with the aggressive behavior shown near the beginning (Ostrov, Gentile, et al., 2006; Ostrov et al., 2013). Thus, even shows that are deemed developmentally appropriate may place children at risk for learning and, ultimately, displaying relational aggression. Future research is needed to examine the role of active parental mediation versus other more passive coviewing parental approaches in these prospective links (see Linder & Werner, 2012).

Developmental Outcomes: Maladaptive Correlates A significant question regarding the development of relational aggression is whether these behaviors are typical and normative, or whether they portend risk for the development of

psychopathology (e.g., Underwood et al., 2001). Consistent with the suggestion that relational aggression is problematic and reflects developmental risk, an extensive literature has demonstrated that such conduct is associated with maladaptive correlates both concurrently and over time. Although relational aggression occasionally is conceptualized as the outcome of other adjustment problems (e.g., Ostrov, 2010; Zimmer-Gembeck et al., 2005) or psychiatric disorders (e.g., Belden, Gaffrey, & Luby, 2012), it is more frequently hypothesized to serve as a precursor or risk factor for psychopathology across development. Prior reviews (see Crick & Zahn-Waxler, 2003; Zahn-Waxler, Crick, Shirtcliff, & Woods, 2006) have demonstrated the theoretical and empirical support for prospective links between relational aggression and maladaptive outcomes. The review in this chapter provides a discussion of some of the most well-documented correlates of relational aggression, with particular attention to recent literature.

Internalizing Problems A number of theorists have suggested that engagement in relational aggression will promote the developmental of internalizing symptoms, including depressive symptoms, anxiety, withdrawn behavior, and somatic complaints. Relational aggression may serve as a risk factor for internalizing problems because these behaviors interfere with successful negotiation of key developmental tasks, such as the establishment of positive peer relationships (e.g., Kamper & Ostrov, 2013; Murray-Close et al., 2007). In addition, relational aggression may be experienced as interpersonally stressful for both perpetrators and victims and may reflect maladaptive beliefs about the self and relationships; these factors, in turn, are often conceptualized as developmental vulnerabilities for depression (e.g., Rudolph et al., 2000). In fact, studies spanning developmental periods have documented positive associations between relational aggression and internalizing problems (e.g., Fite, Stoppelbein, Greening, & Preddy, 2011). Importantly, several of these studies controlled for levels of physical aggression (e.g., Crick, Casas, & Mosher, 1997), indicating that relational aggression may provide unique information regarding these adjustment difficulties. In fact, Underwood, Beron, and Rosen (2011) recently found that high trajectories of social, but not physical, aggression across third through seventh grades were uniquely related to internalizing problems at age 14. Additional longitudinal work suggests that relational aggression serves as a risk factor for the development of internalizing problems. For example, relational aggression has been found to predict increases in internalizing symptoms over time for school-age children (Crick, Ostrov, & Werner, 2006). Moreover, relational aggression and internalizing problems have been found to track together in a dynamic fashion over time during early adolescence (Murray-Close et al., 2007), and some evidence indicates that these associations emerge across cultural contexts (e.g., in Chinese culture; Kawabata et al., 2012). Finally, in a longitudinal study of youth in early adolescence, peer-reported relational aggression was associated with increases in depressive-anxious symptoms (Ellis, Crooks, & Wolfe, 2009). In an effort to isolate the specific internalizing outcomes associated with relational aggression, researchers have also examined relations with narrow-band symptoms of depression and anxiety. In an early investigation in this area, Crick and Grotpeter (1995) demonstrated that relationally aggressive

school-age children (i.e., 1 standard deviation above the sample mean) exhibited higher levels of depressive symptoms (as indexed by the Child Depression Inventory) than their nonaggressive peers. Consistent with the hypothesized role of stress in the emergence of internalizing pathology among relational aggressors, in one recent study, relational aggression predicted heightened levels of depressive symptoms only among youth who frequently ruminated about peer stressors (e.g., being targeted by peer aggression; Mathieson, KlimesDougan, & Crick, 2014). Interestingly, some evidence suggests that the association between relational aggression and depressive symptoms is stronger for girls than for boys. For instance, Crick et al. (1997) reported that relational aggression was uniquely (i.e., controlling for physical aggression) associated with depressed affect for girls, but not boys, during early childhood. In one of the largest longitudinal studies on the topic to date, Spieker et al. (2012) documented that relational aggression during middle childhood was associated with more depressive symptoms at age 15, but only for girls (although see Underwood et al., 2011). These gender differences may reflect the important role of interpersonal stress in the development of depressive symptoms for girls (see Ostrov & Godleski, 2010). Past research has also examined associations between relational aggression and symptoms of anxiety. Anxious youth may use relational aggression to divert negative peer attention to others (Marsee, Weems, & Taylor, 2008) or in response to perceived negative evaluations (Gros et al., 2010). In fact, in a psychiatric inpatient sample of 4- to 12-year-olds, relational aggression was associated with heightened levels of anxiety (Becker, Luebbe, Stoppelbein, Greening, & Fite, 2012). In another study, reactive relational aggression predicted increased levels of anxiety in a sample of 6- to 17-year-olds (Marsee et al., 2008). Studies in emerging adulthood have also documented positive associations between relational aggression and multiple types of anxiety (Gros et al., 2010), including social anxiety (Storch, Bagner, Geffken, & Baumeister, 2004). Given the suggestion that relational aggression may be used as a strategy to deal with anxiety (e.g., Marsee et al., 2008), future longitudinal work is necessary to explore the direction of effects between relational aggression and symptoms of anxiety. Moreover, since similar underlying processes (e.g., dysregulated emotional responses) may portend risk for both anxiety and relationally aggressive conduct (Marsee et al., 2008), it will be important for future research to examine whether symptoms of anxiety and relational aggression share common developmental pathways.

Externalizing Problems Relational aggression may be related to indices of externalizing pathology for several reasons (Crick & Zahn-Waxler, 2003). On one hand, several researchers have hypothesized that relational aggression will serve as a risk factor for the development of externalizing symptoms, including attention problems, conduct disorder symptoms, and oppositional defiant disorder symptoms (e.g., Crick, Ostrov, & Werner, 2006; Prinstein, Boergers, & Vernberg, 2001). From this perspective, engagement in relational aggression may place youth on a developmental trajectory toward adjustment problems, perhaps in part because such conduct interferes with successful functioning across a variety of domains (e.g., peer relationships; internalizing problems; see Crick, Ostrov, & Werner, 2006). For instance, relationally aggressive youth may

select into friendships with similarly aggressive peers; in turn, these peers may socialize youth to increasingly engage in negative behaviors (Sijtsema, Ojanen, et al., 2010). Alternatively, it has been posited that relational aggression is part of a broader antisocial behavior or externalizing construct (see Burt, Donnellan, & Tackett, 2012; Tackett & Ostrov, 2010; Tackett et al., 2009) and serves as a marker of conduct and other externalizing problems (e.g., Tiet, Wasserman, Loeber, McReynolds, & Miller, 2001). From this perspective, relational aggression may be associated with other indices of externalizing symptomatology because they are related manifestations of a similar underlying pathology. In fact, an emerging body of research indicates that relational aggression is positively associated with various indices of externalizing symptoms. For instance, relational aggression has been found to be related to heightened hyperactivity-impulsivity in early childhood (Ostrov & Godleski, 2009) and symptoms of ADHD in middle childhood (Tseng et al., 2012). The links between relational aggression and ADHD symptoms appear robust and have been documented in several cultures (e.g., a Taiwanese sample; Tseng et al., 2012) and in normative as well as clinical samples (e.g., an inpatient psychiatric sample; Becker et al., 2012; although see Belden et al., 2012). For instance, Zalecki and Hinshaw (2004) reported that relational aggression was most prevalent among girls (ages 6 to 12) attending a summer camp who were diagnosed as having ADHD-combined subtype, rather than ADHD-inattentive subtype, or nondiagnosed comparison girls. Relational aggression is also associated with other forms of externalizing pathology; for instance, in a study with adults, relational aggression was elevated among individuals with intermittent explosive disorder (Murray-Close et al., 2010). Developmental psychopathologists have also documented unique associations between relational aggression and conduct problems. Findings from the NICHD SECCYD indicated that relational aggression during middle childhood predicted heightened delinquency and risky behavior (Spieker et al., 2012). Relational aggression was also moderately associated with symptoms of oppositional defiant disorder and conduct disorder in a study of 9- to 17-year-old girls and boys (Keenan, Coyne, & Lahey, 2008). Some evidence indicates that relational aggression serves as a risk factor for later conduct problems; in one study, relational aggression among school-age children was uniquely related to future delinquency over time, controlling for physical aggression (Crick, Ostrov, & Werner, 2006). Several researchers have documented gender differences in these relations; for instance, Prinstein and colleagues (2001) found that relational aggression was associated with externalizing symptoms, but only for girls. However, others have reported stronger associations for boys; for instance, although Spieker et al. (2012) reported that relational aggression was related to risky behavior for both boys and girls, the strength of the association was stronger for boys. Despite these interesting findings, future work is needed to better understand whether relational aggression places youth on a developmental trajectory toward externalizing symptoms and/or whether relational aggression is best conceptualized as a marker or indicator of externalizing difficulties. In their confirmatory factor analyses of disruptive behavior in a large, representative community sample of girls, Loeber and colleagues (2009) found that relational aggression consistently emerged as a unique factor (i.e., along with oppositional behavior/conduct problems, inattention, hyperactivity impulsivity) with both parent and teacher

ratings. These findings suggest that despite overlap in constructs, relational aggression is distinct from existing measures of conduct problems. However, the question remains regarding whether relational aggression is best conceptualized as a subset of antisocial or externalizing behavior problems. Burt et al. (2012) reported that social aggression was uniquely associated with known predictors and correlates of antisocial behavior, suggesting that such conduct may reflect a distinct subtype of antisocial conduct. Moreover, although externalizing problems are often conceptualized as developmental outcomes of relational aggression, effects are likely bidirectional in nature; in fact, Belden et al. (2012) reported that youth with preschool-onset disruptive disorders were more than five times as likely to be identified as relational aggressors 2 years later. Similarly, a 4.5-year follow-up of Zalecki and Hinshaw's (2004) summer camp study revealed that multiple informants rated girls with ADHD to have higher relational aggression scores relative to comparison girls, with moderate to large effect sizes (Mikami et al., 2008). One potential interpretation of these bidirectional associations is that relational aggression is a behavioral manifestation of underlying externalizing pathology or antisocial conduct (see Burt et al., 2012); in effect, engagement in externalizing behaviors may encourage conceptually related behaviors, such as relational aggression. We echo other calls to continue to examine how relational aggression findings are similar or different from those of “pathological externalizing behaviors” (Tackett et al., 2009, p. 731).

Peer Problems Several researchers have argued that relational aggression will increase risk for peer problems and other social concerns (e.g., Preddy & Fite, 2012), which may in turn interfere with the successful negotiation of stage-salient developmental tasks in middle childhood and adolescence (Sroufe, Egeland, & Carlson, 1999). Scholars have posited that because children view relationally aggressive behavior as harmful and distressing (Crick, Bigbee, & Howes, 1996), victims and other peers may dislike aggressors. Crick and Grotpeter (1995) were the first to show that relationally aggressive children (1 standard deviation above the sample mean) were more rejected by their peers than nonaggressive children. Crick et al. (1997) also revealed positive associations between peer-reported relational aggression and peer-reported peer rejection when controlling for physical aggression. Crick, Ostrov, Burr, et al. (2006) documented that observations of relational aggression predicted future teacher-reported peer rejection, addressing potential concerns of shared method variance in earlier work in the area. Additionally, Tseng and colleagues (2013) reported that relational aggression was associated with increases in peer rejection across the fifth grade in a sample of Taiwanese youth. Studies examining subtypes of relational aggression have found that proactive relational aggression is uniquely associated with decreases in peer rejection and relational victimization over time, whereas reactive relational aggression is uniquely associated with increases in peer rejection and relational victimization over time (Ostrov et al., 2013, 2014). Collectively, these findings highlight the possibility that some functions of relational aggression may be used to achieve positive status in the peer group (see the “Developmental Outcomes: Potential Positive Correlates” section).

Peer rejection resulting from relational aggression may further compromise perpetrators' ability to form high-quality peer relationships, ultimately leading to feelings of loneliness and psychological distress (Ostrov & Godleski, 2013) or resulting in victimization by peers (Ostrov, 2008). In fact, past work has shown that relational aggression is associated with feelings of loneliness (Soensens, Vansteenkiste, Goossens, Duriez, & Niemiec, 2008). Additionally, consistent with a sequential social process model, Ostrov (2008) found that relational aggression was associated with peer rejection, which, in turn, placed young children at risk for future relational victimization. Testing a similar model using data from the NICHD SECCYD, Ostrov and Godleski (2013) provided evidence for an indirect social process model such that the direct link between relational aggression in third grade and relational victimization in sixth grade was mediated by loneliness in fifth grade. These findings highlight the relatively complex mechanisms linking relational aggression and negative peer outcomes. To date, the majority of research regarding how relational aggression relates to problems with peers has focused on difficulties within the larger peer group (e.g., rejection or victimization). Fewer studies have been conducted that examine the implications of relationally aggressive behaviors for adjustment in the context of close friendships and romantic relationships. However, the consequences of relational aggression may differ across relational contexts; in fact, Goldstein and Tisak (2004) reported that adolescents' and young adults' expectations regarding the outcomes of relational aggression (e.g., whether the relationship partner would want to continue the relationship if he or she used relational aggression; likelihood of retaliation) differed by relational context. Additionally, Murray-Close et al. (2010) reported that risk factors for relational aggression (e.g., HAB for relational provocations, anger) were positively associated with romantic relational aggression, even when controlling for peerbased relational aggression, raising the possibility that individuals who engage in relational aggression across relational contexts may be at particularly high risk for the development of psychopathology. Moreover, there is good reason to believe that relationally aggressive behaviors in close relationships will be particularly harmful as such conduct violates the common expectation of trust and intimacy (see Goldstein & Tisak, 2004). In fact, several researchers have found that relational aggression is associated with difficulties in dyadic relationships with peers. Interestingly, evidence indicates that relationally aggressive youth often do have close, dyadic friendships. For instance, Rys and Bear (1997) found that relationally aggressive children were just as likely to have a friend as their nonaggressive peers (see also Burr, Ostrov, Jansen, Cullerton-Sen, & Crick, 2005; although see Johnson & Foster, 2005). However, although relationally aggressive youth often have close friends, these negative social behaviors may lead to poor-quality relationships over time. In fact, relationally aggressive youth are more likely than their nonaggressive peers to exhibit difficulties within friendships, including high levels of jealousy and a tendency to use relational aggression against close friends (Grotpeter & Crick, 1996). Banny, Heilbron, Ames, and Prinstein (2011) provided longitudinal evidence that relationally aggressive talk in the context of a friendship interaction (e.g., gossip about others) was associated with increases in negative relationship quality across 6 months. Similarly, Kamper and Ostrov (2013) found that the perpetration of relational aggression in the fifth grade predicted lower friendship quality 1 year later, even

when controlling for engagement in physical aggression. Romantic relational aggression is also associated with negative relationship qualities. For instance, Linder et al. (2002) reported that the use of relational aggression against romantic partners was positively associated with feelings of frustration, ambivalence, jealousy, and anxious clinging and negatively associated with trust in the relationship. Interestingly, this study also found evidence of cross-contextual continuity, indicating that poor relationships with peers and parents may play an important role in the development and maintenance of relational aggression in romantic relationships (Linder et al., 2002). Taken together, then, evidence indicates that relational aggression is associated with impaired peer functioning, at both the dyadic and the peer group levels. However, several factors may be important to consider in research regarding relational aggression and peer outcomes. For instance, findings may differ by gender; in fact, a study in middle childhood found that physical aggression was predictive of social problems for boys, whereas relational aggression was predictive of social problems for girls (Preddy & Fite, 2012). In a recent longitudinal study, relational aggression predicted increases in relational victimization over the course of 8 months in adolescent girls but not boys (Zimmer-Gembeck & Duffy, 2014). Crick et al. (2009) suggested that relational aggression may be more strongly related to poor peer outcomes for girls because these behaviors disrupt the interpersonal goals central to female peer groups. There may also be cultural differences in the strength of the association between relational aggression and negative peer outcomes. Peer problems such as peer rejection have been documented in youth who engage in high levels of relational aggression in numerous cultures (e.g., a large sample of Italian preschoolers; Nelson, Robinson, Hart, Albano, & Marshall, 2010). However, relational aggression may be especially disruptive to peer functioning in some cultures; for instance, relational aggression may interfere with the social expectations typical of collectivistic cultures (Kawabata et al., 2012) and thus may be more strongly related to rejection by peers in this context (see the “Cultural Perspectives” section). Associations between relational aggression and peer outcomes may also be moderated by individual differences in functioning in other domains; for example, past research has shown that relational aggression was associated with conflictual friendships (a marker of negative friendship quality) for children who were rejected by their peers but not for those who were perceived as popular (Rose, Swenson, & Carlson, 2004). In another study, relationally aggressive girls with high levels of sensitivity to relational threats (e.g., rejection sensitivity, fear of negative evaluation, and intimacy avoidance) were at particularly high risk for concurrent and future relational victimization (Zimmer-Gembeck & Duffy, 2014). Importantly, although impaired peer functioning is often conceptualized as an outcome of relationally aggressive behavior, it is likely that associations are bidirectional in nature (see the “Risk Factors: Social Processes” section).

Substance Use and Abuse Relational aggression may place youth on trajectories toward risky behaviors such as substance use (e.g., Kamper & Ostrov, 2013; Spieker et al., 2012). In fact, emerging research findings have documented positive associations between relational aggression and substance

use (e.g., Burt et al., 2012; Storch et al., 2004). For instance, self-reported relational aggression has been significantly correlated with self-reported cigarette use, alcohol use, marijuana use, and advanced alcohol use in a typically developing sample of eighth graders (Sullivan et al., 2006). Some researchers have documented gender differences in these associations; after controlling for initial levels of physical aggression and baseline drug use, Skara et al. (2008) found that relational aggression predicted cigarette use and marijuana use for females but not for males (see also Burt et al., 2012). Relational aggression within the context of romantic relationships has also been shown to be associated with alcohol and drug use (Bagner, Storch, & Preston, 2007). Several peer processes may underlie the relation between relational aggression and substance use problems. For example, relationally aggressive youth may increasingly associate with aggressive and delinquent peers (e.g., Sijtsema, Ojanen, et al., 2010; Werner & Crick, 2004), resulting in socialization toward delinquent acts, such as early substance use and abuse. Moreover, peer difficulties resulting from relational aggression (e.g., rejection, victimization) may promote substance use. That is, relational aggressors may engage in substance use as a maladaptive coping mechanism to deal with their peer difficulties, including being targeted by peer victimization (see Sullivan et al., 2006). Kamper and Ostrov (2013) recently documented that negative friendship quality mediated the longitudinal association between childhood relational aggression and adolescent risky behaviors, including substance use. Future research is needed to further delineate the specific developmental mechanisms underlying substance use among relationally aggressive youth, with particular attention to the role of peer difficulties, such as victimization and poor friendship quality.

Personality Pathology In recent years, several associations between relational aggression and dimensions of personality pathology have been posited and tested (see Cicchetti & Crick, 2009; Tackett & Ostrov, 2010). Although relational aggression has been shown to be associated with several forms of personality pathology including antisocial personality disorder symptoms (e.g., Ostrov & Houston, 2008; Werner & Crick, 1999), the focus of this review is on borderline personality disorder (BPD) features, narcissism, and psychopathy. With respect to symptoms of borderline pathology, Geiger and Crick (2001) conducted a content analysis of the fourth edition of the Diagnostic and Statistical Manual of Mental Disorders (DSM-IV) criteria for diagnosis of personality disorders, and five childhood indicators of BPD were identified. These included: (1) hostile/paranoid view of the world; (2) intense, unstable, inappropriate affect; (3) overly enmeshed and close relationships; (4) impulsivity; and (5) lack of a coherent sense of self. Difficulties within close interpersonal relationships, chronic anger, impulsivity, and a hostile worldview (i.e., HAB) are some of the core features that are shared by both relational aggression and BPD (Crick, Murray-Close, & Woods, 2005; Ostrov et al., 2011). From this perspective, stable behavioral patterns in childhood, including engagement in relational aggression, may reflect risk for the development of borderline symptoms (Crick et al., 2005).

Consistent with this hypothesis, Crick and colleagues (2005) provided evidence indicating that relational aggression may be a unique developmental precursor to and predictor of borderline personality features in children. Subsequent research has supported a conceptual model in which relational aggression is a developmental antecedent of BPD features (e.g., Ostrov & Houston, 2008; Stepp, Pilkonis, Hipwell, Loeber, & Stouthamer-Loeber, 2010). For instance, Vaillancourt et al. (2014) recently reported that relational aggression at age 12 predicted higher levels of borderline symptoms at age 14. Research has further documented prospective associations between social aggression and BPD features (Underwood et al., 2011). Werner and Crick (1999) were the first to empirically document that relational aggression was associated with heightened BPD features in emerging adults. In recent years, this association has been replicated several times (e.g., Ostrov & Houston, 2008; Ostrov et al., 2011), although Burnette and Reppucci (2009) failed to find an association between relational aggression and borderline symptoms in a sample of incarcerated adolescent girls. In one recent longitudinal study of the association between forms of aggression and borderline pathology in a Russian sample, preschool aggression was not associated with later borderline features. However, among males and females, borderline features were concurrently associated with both physical and relational aggression in adolescence (Nelson, Coyne, Swanson, Hart, & Olsen, 2014). Narcissistic personality traits may also be associated with relational forms of aggression. Narcissistic individuals by definition have an enhanced sense of self and status relative to peers; moreover, these individuals desire power and control over others and often exploit and manipulate their peers (Barry, Grafeman, Adler, & Pickard, 2007). This goal for social dominance over peers may promote the use of aggression among individuals with narcissistic traits (Barry et al., 2007). In addition, narcissistic youth may engage in aggression when their inflated self-views are threatened (Bukowski, Schwartzman, Santo, Bagwell, & Adams, 2009). Consistent with these hypotheses, several researchers have reported positive associations between relational forms of aggression and trait narcissism (Burt et al., 2012; Golmaryami & Barry, 2010; Lau & Marsee, 2013; Ojanen, Findley, & Fuller, 2012) and in particular with maladaptive subtypes of narcissism (i.e., entitlement, exhibitionism, and exploitativeness; Barry, Pickard, & Ansel, 2009). Associations between narcissistic traits and relational aggression have also been documented in non-Western cultures (e.g., Japan; Onishi et al., 2012). Moreover, although a large body of work has documented heightened levels of physical forms of aggression among individuals with narcissistic traits, Bukowski et al. (2009) reported that narcissism was more strongly related to relational than physical forms of aggression. Finally, psychopathy, a maladaptive personality disorder that includes traits such as lack of remorse, superficial charm, blunted or shallow affect, manipulation of others, and egocentricity (J. D. Miller & Lynam, 2003), has been found to be associated with relational forms of aggression. As discussed in the section titled “Risk Factors: Cognitive and Emotional Processes,” impaired emotional responses may be a mechanism linking psychopathic traits and relationally aggressive conduct. Initial studies showed links between the two-factor model (i.e., impulsive antisociality and fearless dominance) and relational aggression (e.g., Ostrov & Houston, 2008; Schmeelk, Sylvers, & Lilienfeld, 2008; Warren & Clarbour, 2009; cf. J. D. Miller & Lynam, 2003). Regardless of the methodological approach, robust associations have

been documented between relational aggression (or indirect aggression, Coyne & Thomas, 2008; Vaillancourt & Sunderani, 2011) and measures of psychopathy or CU traits, even after controlling for the variance associated with physical aggression (e.g., Czar et al., 2011; Ostrov & Houston, 2008; Schmeelk et al., 2008). These links have been found with regard to both peer and romantic relational aggression (Coyne et al., 2010; Czar et al., 2011). There are several important directions for future research regarding the association between relational aggression and personality pathology. For instance, some evidence has emerged indicating that distinct personality traits may interact in the prediction of relational aggression (e.g., narcissism and CU traits or measures of psychopathy; Kerig & Stellwagen, 2010). Additionally, personality pathology (e.g., borderline pathology; narcissistic traits) may be more strongly associated with relational aggression in females than in males (Banny, Tseng, Murray-Close, Pitula, & Crick, 2014; Marsee et al., 2005). In fact, in one recent longitudinal study, the indirect effect linking relational aggression at age 10 with borderline symptoms at age 14 via relational aggression at age 12 was significant for girls only; these findings suggest that stable levels of relational aggression may be a stronger risk factor for borderline pathology in females as compared to males (Vaillancourt et al., 2014). Relations between forms of aggression and personality pathology may also differ depending on the function of the aggression (see Banny et al., 2014; Crapanzano, Frick, & Terranova, 2010; Ostrov & Houston, 2008). In one study, proactive but not reactive relational aggression significantly and uniquely predicted CU traits in a sample of adolescent detained girls (Marsee & Frick, 2007). Other work has highlighted mechanisms that might underlie the association between personality traits and relational aggression; for instance, some evidence indicates that the direct association between narcissism and relational aggression is mediated by Machiavellianism (i.e., a tendency to manipulate others and to exploit them for one's own goals; Kerig & Stellwagen, 2010). These findings highlight the need for additional longitudinal research exploring associations between personality pathology and relational aggression, with a particular emphasis on underlying mechanisms and moderating factors. Finally, relational aggression may be conceptualized as both a predictor (e.g., Marsee & Frick, 2007) as well as an outcome (e.g., Onishi et al., 2012) of personality pathology. For instance, Bukowski et al. (2009) argued that narcissistic youth may use relational aggression to defend their inflated self-views. Narcissistic traits may also serve as a risk factor for involvement in relational aggression because they interfere with appropriate emotional responses regarding aggression (e.g., guilt; Onishi et al., 2012). Banny and colleagues (2014) posited that core features of borderline pathology, such as affective instability, may lead to the use of relational aggression; consistent with this suggestion, these authors reported that borderline features were longitudinally associated with increased relational aggression in girls. Some evidence also emerged for the alternative direction of effects, such that earlier relational aggression was associated with the development of borderline pathology; however, whereas reactive relational aggression was marginally associated with increases in borderline features over time, proactive relational aggression was related to decreases in these symptoms (Banny et al., 2014). As with studies of externalizing behavior problems, it will be important for future research to evaluate the extent to which relational aggression can be conceptualized as a

behavioral manifestation of personality pathology. Such work would help clarify the mechanisms underlying the likely dynamic bidirectional associations between relational aggression and personality pathology across development.

Developmental Outcomes: Potential Positive Correlates In considering correlates of relational aggression or similar constructs (indirect or social), one of the most intriguing and controversial foci is the assessment of what may be viewed as positive correlates. In particular, studies have sought to determine whether relational aggression may prove to be a useful tool in securing significant social status (termed hierarchy ascension; Hawley et al., 2007). Traditionally, aggression has been considered to align with peer difficulties such as rejection, which may in turn provoke a host of negative outcomes including more aggressive behavior (see the “Developmental Outcomes: Maladaptive Correlates” section). For example, Murray-Close and Ostrov (2009) found that children who were socially excluded tended to react to the stress of their situation by becoming more relationally aggressive over time. However, in recent years, the idea that relational aggression is uniformly related to maladaptive outcomes has been challenged on both theoretical and empirical grounds. Researchers have instead argued that some aggressive youth exhibit social competence as well as high aggression, and they hold enviable social status positions in the peer group (i.e., are considered popular, socially central, and socially dominant). Evidence of this co-occurrence of aggressive behavior and positive positions in the peer group is apparent in early studies of physical aggression and dominance relations among children (Strayer & Strayer, 1978). The research has rapidly expanded over the past decade to include relational (social, indirect) aggression and to consider how these correlations might vary by child gender. Importantly, aggression itself does not predict status so much as how effectively the child makes use of it to achieve social goals (Prinstein & Cillessen, 2003; Walcott, Upton, Bolen, & Brown, 2008), suggesting that relational aggression may be most likely to translate into popularity among youth who combine these negative behaviors with high social skills or prosocial behavior. In fact, Puckett, Aikins, and Cillessen (2008) found that perceived popularity was associated with social self-efficacy, leadership, cooperation, and peer sociability, highlighting the importance of positive social skills in establishing a position of high status in the peer group. Notably, there is considerable overlap between membership in the perceived popular category and the controversial sociometric status group (Prinstein & Cillessen, 2003). Controversial-status children are frequently selected by peers for both liked and disliked peer nomination items,, and studies of this group point to their high levels of aggression as well. For example, controversial preschool children are just as aggressive as their rejected counterparts, but engagement in average levels of sociability buffers against fullscale rejection by peers, both in the United States (Nelson, Robinson, & Hart, 2005) and in Italy (Nelson et al., 2010). These findings suggest that youth who exhibit a mix of aggressive and socially competent behaviors may not experience significant problems in the peer group and, in some cases, may actually benefit from high-status positions. Moreover, many children already appear to be proficient at this mix of behavior by preschool (Nelson et al., 2005).

A related perspective draws from evolutionary theory, highlighting the role of aggression in securing access to important resources and promoting one's own self-interest (Hawley, 2003a; Hawley, Little, & Card, 2008). Specifically, Hawley (2003a) has proposed that some youth are bistrategic controllers, opting for a combination of prosocial and coercive control tactics in their bid to control resources. For these youth, significant social intelligence and savvy appear to be central to effective engagement in relational aggression (Andreou, 2006). This group is similar to perceived popular or controversial youth in that relational aggression is combined with positive social behaviors and is used effectively to gain access to social resources (e.g., a socially central position in the peer group; Hawley, 2003a; Hawley et al., 2008). An impressive number of studies have documented the positive association between relational aggression and high social status. Particularly influential are the studies that have assessed the interplay of these factors over the course of time. Cillessen and Mayeux (2004) have shown that as children move into adolescence, relational aggression becomes increasingly associated with high perceived popularity but low social preference. Relational aggression and perceived popularity are particularly connected for girls. Accordingly, there are ample reinforcements (social benefits) for those who become proficient in their use of relational aggression in the peer group. Rose, Swenson, and Waller (2004) additionally found that for older girls, the positive relations between relational aggression and perceived popularity were bidirectional in nature. Many of these connections appear to be relevant not only for Caucasian U.S. children but also for ethnic minorities (Waasdorp, Baker, Paskewich, & Leff, 2013). Similar associations are found in a number of samples outside of the United States and Canada, including Greek (Andreou, 2006), Belgian (Kuppens, Grietens, Onghena, Michiels, & Subramanian, 2008) and Italian (Nelson et al., 2010) samples. However, there are findings which indicate cultural differences as well (see the “Cultural Perspectives” section), perhaps reflecting cultural variability in the tolerance of aggressive behavior. Relational aggression appears to promote high status across several relational contexts; for instance, Pellegrini and Long (2003) established that, in middle school, relational aggression is connected to dating popularity, particularly among girls. Being relationally aggressive does not appear to hamper friendship formation for popular aggressive youth either. Friends tend to be similar in their levels of aggression and popularity, creating further insularity from realworld consequences of their aggression toward others (Rose, Swenson, & Carlson, 2004; Peters et al., 2010). The friendships of bistrategic controllers are characterized by positive qualities, such as high intimacy and fun, although they also tend to include high levels of conflict (including overt and relational aggression; Hawley et al., 2007). Some longitudinal evidence also indicates that relational aggression is associated with increases in positive friendship qualities across time, although effects for negative relationship qualities were observed as well (Banny et al., 2011). Banny and colleagues (2011) suggested that relationally aggressive behaviors are one way that youth may establish feelings of closeness, intimacy, and alliance among close friends. As peer censure and difficulties are one of the primary mechanisms thought to account for associations between relational aggression and poor adjustment outcomes (e.g., depressive symptoms; see the “Developmental Outcomes: Maladaptive Correlates” section), popular relationally aggressive youth may not exhibit many

of the problems typically associated with such conduct. Consistent with this suggestion, Rose and Swenson (2009) found that perceived popularity buffered relationally aggressive youth from internalizing problems. Leadbeater, Boone, Sangster, and Mathieson (2006) have cautioned against the impression, however, that aggressive yet popular youth do not pay a price for their hostility toward others since aggression tends to invite retaliation and stress in peer relationships. Based on her work with students in middle childhood, sociologist Donna Eder (1985) described a “cycle of popularity” whereby views of popular girls moved from being positive to very negative (see the “Developmental Outcomes: Maladaptive Correlates” section). Underwood (2004) also noted that popular girls consistently struggle with a balancing act. They want to avoid a reputation for meanness, and nonverbal forms of social aggression (glares, eye rolls, and the silent treatment) may seem the least toxic approach to aggressively defending their status. The irony is that, in doing so, they engage in behaviors that may earn them the reputation of being stuck up, a label they dread (Merten, 1997). What drives the desire for social dominance may also be problematic. Ojanen et al. (2012) provided evidence that narcissistic personality is associated with dominance goals, which are in turn predictive of relational and physical aggression. More important, there is evidence of long-term dysfunction for many popular youth. Perceived popularity predicts increases in general externalizing problems over the course of several years (Sandstrom & Cillessen, 2006), which could place popular youth on a trajectory illsuited for successful relationships beyond the peer group. Indeed, Sandstrom and Cillessen (2010) have demonstrated how levels of perceived popularity in high school may interact with engagement in relational aggression to predict how students will fare in their emergingadulthood years. In particular, greater high school popularity predicted more engagement in risk behaviors post–high school, even after controlling for prior risk behaviors. Boys who were not popular but were highly relationally aggressive in high school had pronounced adjustment problems after high school. For girls, being highly relationally aggressive in high school was associated with less post–high school depression but higher workplace victimization. Additional longitudinal study is needed to fully understand the long-term consequences of being aggressive yet popular. There are several important directions for future research on the association between relational aggression and positive peer outcomes. For instance, researchers should consider whether contextual factors, such as classroom dynamics, moderate the extent to which relationally aggressive youth are able to garner positive peer regard. In one recent study, Moore, Shoulberg, and Murray-Close (2012) reported that relationally aggressive girls were more likely to be popular, and less likely to be rejected, if their teachers liked them. Popularity with teachers may also enhance these children's ability to aggress with relative impunity, as teachers may not look for or see aggressive behavior as readily in children they generally consider to be socially competent (see Hawley, 2003b). The association between relational aggression and high status may also depend on additional classroom factors, such as levels of status hierarchy (Garandeau, Ahn, & Rodkin, 2011).

Additionally, as relational aggression is often conceptualized as a strategy that youth use to achieve status and power with peers, it will be important for researchers to examine whether individual differences in social status motivations (e.g., desires to be well liked or popular in the peer group) are associated with relationally aggressive conduct. In fact, Li and Wright (2014) found that self-reported popularity goals predicted higher levels of self-reported, but not peer-reported, relational aggression in a sample of middle schoolers. Dawes and Xi (2014) found that self-reported popularity goals were associated with higher levels of peerreported social aggression among popular, but not unpopular, middle schoolers. The authors suggested that popular youth who endorse popularity goals may have the social resources necessary to effectively use social aggression. Other researchers have also documented the important role of actual peer status in the association between popularity goals and relational aggression; for instance, Shoulberg, Sijtsema, and Murray-Close (2011) found that having a reputation for valuing popularity was generally predictive of girls' engagement in relational aggression. The association was strongest, however, for two groups of girls who valued popularity: (a) popular girls who exhibited blunted physiological reactivity to social exclusion and (b) unpopular girls (wannabes) who demonstrated heightened physiological reactivity to exclusion. The lower stress reactivity of popular girls may allow them to pivot quickly and use relational aggression to achieve or maintain status. In contrast, the reactivity of unpopular girls may lead to aggressive responses to any perceived social slight, which further compounds their popularity problems. Future research may also benefit from considering whether some manifestations of relational aggression are more strongly associated with positive outcomes than others. Research suggests that popular yet aggressive children adopt a wide range of aggressive behaviors in their bids for social dominance. In the context of social aggression, these children may employ both verbal and nonverbal expressions of relational manipulation. Glares of contempt, rolling of the eyes to signal disgust, and the silent treatment are nonverbal forms of social exclusion that can exert tremendous influence (Underwood, 2004). The verbal/nonverbal distinction is a useful one; when controlling for both overt and verbal social aggression, nonverbal social aggression uniquely accounts for variance in children's social status (Blake et al., 2011). Fanger et al. (2012) investigated various types of exclusionary behavior among preschoolers and argued, however, that not all types may be considered aggressive. For example, a child may seek to keep high-quality interactions going with certain peers by not allowing others to interrupt the flow. Many of the exclusionary behaviors exhibited by preschoolers are relatively indirect (subtle) in nature or focused on simply ignoring a play overture. Better-accepted preschoolers, in particular, frequently employ these forms of exclusionary behavior, and they may not hold hostile intent toward the peers they exclude. These findings highlight the importance of research assessing whether distinct relationally aggressive behaviors have differential associations with positive outcomes, such as popularity (see the “Developmental Manifestations of Relational Aggression” section). Additional research is clearly necessary to explore the significant heterogeneity among relationally aggressive children, including the fact that some youth appear to suffer from rejection whereas others exhibit significant social status benefits as a result of their aggressive conduct.

Cultural Perspectives Studies of relational aggression and similar constructs (e.g., indirect and social aggression) have been conducted across a number of diverse cultures. Similar to the study of physical aggression, a major impetus of this research is to gauge cross-cultural similarity or difference in the quantity, quality, and meaning of relationally manipulative behaviors. This work has provided significant insight into the development of relational aggression in context and has highlighted both challenges and opportunities for relational aggression researchers. More specifically, cross-cultural research underscores the need to develop culturally sensitive tools for assessing relational aggression, to consider the contextual processes that might encourage engagement in relational aggression, and to investigate potentially distinct predictors and developmental outcomes associated with relationally aggressive conduct. A significant challenge in cross-cultural research is the development of measures that are appropriate for use across cultural contexts. In many of the existing studies of relational aggression conducted outside of the United States and Canada (the source of the majority of relational aggression research to date), existing measures are translated for use in other languages, and their psychometric properties are assessed (e.g., Kawabata et al., 2012; Lansford et al., 2012). The standard procedure of forward- and back-translation helps to ensure the linguistic and conceptual equivalence of the measures across the languages of interest. In regard to these goals, measures of relational aggression developed in Western countries appear to work rather well in diverse cultures. For instance, teacher-report measures developed in the United States exhibited high internal consistency when administered to a sample in China (e.g., Kawabata et al., 2012). Researchers have also documented configural invariance (similarity of factor structures) in measures of physical and relational aggression across cultures, suggesting that relational aggression is distinct from physical aggression and that similar behaviors (e.g., exclusion) underlie relational aggression in diverse settings (e.g., Lansford et al., 2012). Although an emerging body of research suggests that Western measures can be translated and used effectively in other cultures, such an approach may obscure the unique manifestations of relational aggression that may exist across cultures. In fact, some researchers have argued that relational aggression may be expressed in distinct ways in different cultures; for instance, subtle relationally aggressive behaviors (e.g., threatening to tell teachers bad things about a peer) may be more typical in collectivistic than in individualistic societies (e.g., Kawabata et al., 2012). Thus, extant measures may need to be adapted to best capture the various types of relationally aggressive behaviors that are most common in distinct cultural contexts. Culturally unique manifestations of relational aggression may not lend themselves to cross-cultural comparison; however, they may play a significant role in children's adjustment. Accordingly, observational or qualitative studies of aggression may provide significant insight regarding culturally sensitive approaches to assessing relational aggression. A significant body of research has examined relational aggression in countries other than the United States and Canada. For instance, relational aggression and/or victimization have been successfully measured in studies conducted with youth in Colombia (Velásquez, Santo,

Saldarriaga, López, & Bukowksi, 2010), Italy (Nelson et al., 2010; Tomada & Schneider, 1997), China (e.g., Li et al., 2011; Nelson et al., 2006), Russia (Hart et al., 1998), Indonesia (French, Jansen, & Pidada, 2002), Japan (e.g., Kawabata et al., 2012), Greece (Andreou, 2006), Belgium (Kuppens et al., 2008), India (Bowker et al., 2012), and Australia (e.g., Hemphill et al., 2010; Pronk & Zimmer-Gembeck, 2010). Studies of indirect aggression have also included samples from diverse countries, such as the United Kingdom (Coyne, Archer, & Eslea, 2006), Germany (Möller & Krahé, 2009), Australia (Owens, Shute, & Slee, 2000), Finland (Österman et al., 1994, 1998), Poland (Österman et al., 1994, 1998), Israel (Österman et al., 1998), and Italy (Österman et al., 1998). These studies, with participants ranging in age from preschool to adolescence, have documented that relational aggression occurs with some frequency across a wide span of cultural settings. Moreover, evidence indicates that youth spontaneously identify relationally aggressive behaviors as common in peer interactions in varied cultural contexts. For example, French et al. (2002) conducted a study in which U.S. and Indonesian children were asked to describe disliked peers, and coders documented their references to physical, verbal, and relational aggression. Findings showed that youth in both countries frequency cited relationship manipulation (e.g., ignoring peers), social ostracism, and rumor-spreading as behaviors that elicited their dislike. Moreover, girls in both countries were more likely than boys to identify relationally aggressive behaviors in their descriptions. These findings suggest that relational aggression is a common behavior across two cultures that otherwise differ considerably in regard to gender roles, attitudes toward aggression, and along the individualism/collectivism dimension. A critical question is whether relational aggression is uniquely tied to cultural elements, such as a society's place on the collectivism/individualism continuum. Of significant import, therefore, are studies of relational or indirect aggression that directly compare data from diverse cultures. Österman and colleagues (1994) hypothesized that children will select aggressive behaviors based on their risks versus benefits. It is likely that both the normative nature of aggression subtypes and the likelihood that youth will be punished for engaging in these behaviors vary significantly across cultural contexts; as a result, cultural differences in the benefit-to-risk ratio are expected. For instance, given the focus on interpersonal harmony in collectivistic cultures, relational aggression may be less tolerated and thus less common in these contexts (e.g., Forbes, Zhang, Doroszewicz, & Haas, 2009). Consistent with this suggestion, Forbes et al. (2009) documented that engagement in indirect aggression was more common in individualistic (United States) as compared to collectivistic societies (China) or hybrid societies with elements of both individualism and collectivism (Poland). Similarly, Österman et al. (1994) compared levels of physical, verbal, and indirect aggression for children in Finland (Finnish or Swedish), the United States (Chicago, IL: African American or Caucasian), and Warsaw, Poland. Results revealed significant differences in the frequency of aggressive behavior across countries, such that youth in the United States were especially aggressive. As with gender differences (see the “Toward an Integrated Gender-Linked Model of Aggression” section), however, it may be especially important for cross-cultural researchers to

compare cultural differences in within-group preferences for relationally versus physically aggressive behaviors. For instance, collectivistic cultures may discourage the use of all types of aggression because they interfere with harmonious group functioning; however, when youth from collectivistic societies are aggressive, they may be particularly likely to adopt relationally aggressive rather than physically aggressive behavioral strategies (see Kawabata et al., 2012). Recently, Lansford and colleagues (2012) examined this possibility in their comparison of levels of relational aggression versus physical aggression in nine different countries. Results indicated that children reported being more relationally than physically aggressive in China, India, and Thailand. In contrast, Jordanian and Kenyan children reported more engagement in physical than relational aggression. In the remaining four countries (Colombia, Philippines, Sweden, and the United States), there was no significant difference in levels of engagement in these forms of aggression. These findings hint at distinct cultural norms for aggression and underscore the importance of considering cultural differences in withingroup preferences for relational versus physical aggression, in addition to potential crosscultural differences in mean levels of relational aggression. Evidence from several studies also highlights meaningful differences in the use of relational aggression within countries, perhaps reflecting distinct subcultural values and experiences. For instance, Li, Wang, Wang, and Shi (2010) demonstrated that individual differences in endorsement of collectivism or individualism were predictive of Chinese adolescents' engagement in overt and relational aggression. As expected, Chinese adolescents who endorsed more individualism were more aggressive than their collectivism-minded peers. Religious subcultures might also promote moral emotion and thereby reduce youth involvement in aggression. Consistent with this possibility, Landau, Björkqvist, Lagerspetz, Österman, and Gideon (2002) reported that levels of direct and indirect aggression were higher among secular as compared to religious Israeli children and adolescents. Taken together, findings underscore the need for research that examines subcultural variations in the development of relational aggression. Differences in socioeconomic status (SES), urban/rural environment, and religious and familial value systems may all provide significant insights in this regard. Cultural identity may also serve a protective function against the development of relational aggression. Although cultures may vary in their prescriptions for mature social behavior, a positive sense of identification with ancestral culture may provide an important psychological foundation for psychosocial maturity. As a result, cultural identification may promote a more tolerant and less aggressive atmosphere among children and adolescents. Consistent with this suggestion, Flanagan and colleagues (2011) found that aboriginal Naskapi youth in Quebec, Canada, who exhibited a strong cultural identity with their native culture, exhibited less engagement in both relational and physical forms of aggression. In addition, Belgrave et al. (2004) conducted a cultural intervention for African-American girls called the Sisters of Nia, in which relevant cultural values and beliefs (tied to ethnic identity, gender roles, and relational orientation) were strengthened. Findings indicated that the intervention was successful in increasing awareness and adoption of these values and beliefs; moreover, participants in the intervention exhibited marginally lower levels of relational aggression postintervention than those in the comparison group.

A critical question for cross-cultural research is whether the developmental outcomes associated with relational aggression differ across cultural contexts. This perspective highlights the importance of the goodness of fit between cultural norms and engagement in relational aggression. For example, relational aggression may prove more disruptive in collectivist cultures, which may in turn decrease the likelihood that such conduct will garner positive peer regard and increase the risk of poor adjustment outcomes among aggressors. Consistent with this possibility, Butovskaya, Timentschik, and Burkova (2007) found no connection between popularity and verbal, physical, or indirect aggression (based on peer ratings) among Russian adolescents. Moreover, Tseng and colleagues (2013) reported that relational aggression was related to decreases in perceived popularity across the fifth grade among Taiwanese youth. Similarly, Bowker et al. (2012) found that relational aggression was negatively associated with perceived popularity (after accounting for overt aggression) for a sample of adolescents in India. In contrast, overt aggression was related to heightened levels of perceived popularity. These findings suggest that, in India, overt aggression may be the more attractive and socially acceptable way to assert dominance. Additionally, research by Kawabata and colleagues (2010) documented that relational aggression was more strongly associated with symptoms of depression in a Japanese sample than in a contrasting U.S. sample. These findings are consistent with the notion that the stress associated with engagement in relational aggression may be greater in collectivistic, as compared to individualistic, cultures. More generally, findings suggest that children in collectivistic societies may not view relational aggression as favorably as children in individualistic societies; as such, these behaviors may be particularly likely to result in poor developmental outcomes in these contexts. Interestingly, recent research in the United States has documented that relational aggression is more strongly associated with internalizing pathology among Asian American, as compared to European American, youth (Kawabata & Crick, 2013), underscoring the importance of exploring differences in outcomes across racial and ethnic groups within as well as across countries. In summary, a growing body of research has highlighted the salience of relational and indirect aggression across cultural settings. However, additional research is clearly needed to investigate cultural differences in the prevalence of and the developmental outcomes associated with relationally aggressive conduct. It will also be important for future research to explore cultural differences in the manifestation of relational aggression and to develop culturally sensitive tools for assessing these behavioral patterns. Such knowledge will be crucial for researchers developing interventions to reduce relational aggression across cultural contexts. We also anticipate that culturally universal elements will continue to emerge in much of this work. For example, harsh and manipulative parenting may be expected to predict relational aggression across cultural settings (e.g., Nelson et al., 2006). Accordingly, a cultural emphasis need not detract from important commonalities that may be pursued to enhance the lives of children in diverse contexts.

Relational Aggression Interventions

Since the first publication using the term relational aggression (Crick & Grotpeter, 1995), a robust body of research has emerged indicating that both relational aggressors and relational victims are at risk for a number of adjustment difficulties (e.g., peer rejection, internalizing problems, externalizing problems; see the “Developmental Outcomes: Maladaptive Correlates” section). As such, there has been substantial interest among researchers and other professionals in the development of interventions to reduce relational aggression. As others have noted, programs that have been designed to address physical aggression may not be easily adapted for use with relational aggression (Ostrov et al., 2009; Zahn-Waxler et al., 2006), highlighting the need for interventions that specifically target relationally aggressive conduct. An exhaustive review of relational aggression interventions by Leff and colleagues (2010) yielded 21 separate interventions with a focus on relational aggression or related constructs (i.e., indirect aggression, social aggression, social exclusion). Although this total may seem high for an area of research that has systematically been around for only two decades, only 9 of the 21 programs meet the criteria for establishing efficacy as outlined by the Society for Prevention Research (see Flay et al., 2005). However, preliminary findings indicate that several of these programs hold significant promise in efforts to reduce children's involvement in relational aggression (for a review, see Leff et al., 2010). A review of the various programs that have been developed to date is beyond the scope of this chapter; however, a brief overview of exemplar interventions targeting relational aggression at different points in development can be illuminating in terms of both program structure and efficacy findings. Although very few interventions have been developed for the early childhood years, the Early Childhood Friendship Project has shown a great deal of promise (Ostrov et al., 2009). The program is a classroom-based intervention that is designed to reduce relational aggression and relational victimization. The initial intervention (Ostrov et al., 2009) was a 6-week program using observational assessments of relational aggression and victimization in nine randomized intervention classrooms and nine control classrooms. Each week, a different lesson was introduced and reinforced throughout the week. Two of the six lessons focused on relational themes (i.e., relational aggression and relational inclusion). In this study, classrooms were the unit of analysis, and researchers found that effect sizes indicated that the intervention classrooms tended to have greater reductions in both relational aggression and relational victimization. A second illustrative intervention, the Walk away, Ignore, Talk it out, Seek help (WITS) program, was created by Leadbeater, Hoglund, and Woods (2003) and was initially conducted with children in grades K–3. This program focuses specifically on relational victimization and adopts an ecological approach including parents, siblings, and local police in addition to the school-based component. The core of the program utilizes a literacy-based approach that can be readily incorporated into the existing curriculum and lesson plans. For instance, one component of this intervention involves reading books with lessons on social behaviors to participants and then providing them with the opportunity to discuss the story and complete activity sheets. Some evidence supports the efficacy of the WITS intervention; for instance, researchers have found that relational victimization decreased in all of the program schools and in the low-poverty control schools but increased in the high-poverty control schools. The

program has since been expanded through sixth grade, also with encouraging results (Hoglund, Hosan, & Leadbeater, 2012; Woods, Coyle, Hoglund, & Leadbeater, 2007). The Friend to Friend program (Leff et al., 2009) uses a social information-processing approach to target relational aggression among third- through fifth-grade girls in an urban setting. The inner-city context, in addition to the participatory action research framework and its cultural sensitivities, is a particular strength of this program. The program utilizes a socialcognitive retraining model consistent with the Crick and Dodge (1994) reformulated social information-processing model. Specifically, it was designed to help relationally aggressive girls explore and practice the identification of feelings and signs of physiological arousal, calming strategies, interpreting intentions of others, and generating and evaluating alternative behaviors to enact. Empirical evaluations of the efficacy of this program indicated that relationally aggressive girls, randomized to the program, had large decreases in teacherreported relational aggression (Leff et al., 2009). Findings underscore the promise of interventions informed by social information-processing theory in reducing relational aggression. Finally, one recent intervention targeting bullying in Finnish schools, the KiVa program (Kärnä et al., 2011), was developed to address the potential social status benefits of aggressive behavior. More specifically, KiVa builds on the idea that interventions can successfully reduce levels of aggressive behavior by changing the reactions of peers to bullying behaviors (Kärnä et al., 2011). For instance, if bystanders defend victims rather than reinforce displays of aggression, aggressive behavior may no longer serve as a mechanism for youth to gain social power and status. One of the primary components of the KiVa program focuses on helping youth develop skills to support victims of bullying. Recent research highlights the efficacy of the KiVa program in reducing aggressive behavior across developmental periods (e.g., Kärnä et al., 2011), although some evidence indicates that the program is more robustly efficacious in younger as compared to older samples (Kärnä et al., 2013). Importantly, preliminary evidence supports the notion that this program is effective at reducing relational forms of aggression, such as social exclusion, though effect sizes tend to be smaller for these forms of aggression than for physical aggression (Salmivalli, Kärnä, & Poskiparta, 2011). Despite promising findings in this area, research regarding relational aggression interventions is still in its infancy, and several important directions for future research remain. Leff and colleagues' (2010) review of extant relational aggression interventions indicates that the majority of the programs to date have been designed and evaluated for late elementary and/or early middle school children. This is surprising from a developmental standpoint, given that it is likely easier to change behaviors before they become an ingrained part of children's behavioral repertoires. Additionally, the vast majority of these interventions were conducted in a school or after-school setting, highlighting the need for intervention programs that target maladaptive family processes that may contribute to the development of relational aggression (see the “Risk Factors: Social Processes” section). Interestingly, very few of the programs to date have been directed toward relational victims (Leff et al., 2010). Although it is likely that decreases in relational aggression will also benefit the targets of these behaviors, separate programs may be needed to ameliorate the difficulties encountered by victimized youth.

Additionally, research is needed to explore whether different intervention programs are most efficacious across different points in development, for males versus females, or for youth in different cultural settings.

Developmental Psychopathology Perspectives Developmental Tasks A developmental psychopathology perspective advances our understanding of relational aggression in several important ways. First, developmental psychopathologists emphasize the importance of the successful negotiation of key developmental tasks in the emergence of disorder (Cicchetti, 1993). From this perspective, many of the risk factors for relationally aggressive conduct reflect a failure to master salient developmental tasks. As an example, the ability to manage negative emotional reactions increases throughout childhood, and a failure to develop these skills is associated with maladaptive behavioral and emotional patterns (Izard, Youngstrom, Fine, Mostow, & Trentacosta, 2006). Similarly, establishing positive relationships with peers, including garnering acceptance from the peer group, is a key developmental task during middle childhood (Masten & Coatsworth, 1998). As these examples illustrate, both child characteristics (e.g., emotional processes, accurate attributions of others' behaviors) and social risk factors (e.g., peer rejection) for relational aggression can be conceptualized as deviations from typical developmental pathways that reflect a failure to master salient developmental tasks. A developmental psychopathology framework also highlights the possibility that distinct risk factors for relational aggression may emerge across developmental periods. As key tasks differ across developmental stages (Masten & Coatsworth, 1998), it is possible that the factors that are most strongly associated with relational aggression change throughout development. A failure to inhibit angry reactions to frustration, for instance, may be most strongly associated with relational aggression during developmental periods in which mastery of these capacities is salient (e.g., early childhood). In contrast, a failure to establish positive relationships with peers may be most strongly associated with relational aggression when relationships with peers are especially important (e.g., middle childhood and adolescence). However, successful negotiation of key tasks in any given developmental period provides the foundation for later achievements (e.g., the ability to regulate angry reactions likely facilitates the establishment of positive peer relationships; see Cicchetti, 1993); thus, it is critical to understand how the risk factors for relational aggression work together across development to promote maladaptive behavior patterns. A developmental tasks perspective may also provide significant insight regarding the association between relational aggression and maladaptive outcomes. Masten and Coatsworth (1998) argued that engaging in rule-governed conduct, including inhibiting aggressive impulses, is a key developmental task of middle childhood; however, empirical work regarding normative developmental change in relational aggression highlights the possibility that some involvement in relational aggression may be fairly common during middle childhood

and early adolescence. In these developmental contexts, low levels of relational aggression may not portend adjustment difficulties (see Underwood et al., 2011, for an application of this idea to social aggression). In contrast, frequent and high-intensity displays of such conduct may reflect a failure to master this key developmental task and may in part emerge as a result of earlier developmental failures (e.g., an inability to develop appropriate emotion regulation capacities; see Underwood et al., 2011). Thus, a developmental tasks perspective may help researchers identify patterns of relationally aggressive behaviors (e.g., high intensity, frequent, and resulting from dysregulated reactions to stress) that signify particularly high levels of risk for poor adjustment outcomes.

Multilevel Perspectives Consistent with the suggestion by developmental psychopathologists that multilevel perspectives are critical to understanding maladaptive behavior patterns (e.g., Cicchetti, 1993), the development of relational forms of aggression appears to reflect processes ranging from genetic factors to socialization experiences. In recent years, theorists have begun to propose mediation and moderation models that integrate these distinct risk factors. Mediation Models The mechanisms involved in the development of relationally aggressive behavior may occur across multiple levels of analysis, and several lines of research suggest that biological risk factors influence relational aggression via alterations in cognitive, emotional, and social functioning. For instance, researchers have highlighted the need to consider endophenotypes, including psychophysiological and cognitive patterns, in evaluating the processes that link genetic risk with developmental outcomes (Brendgen, 2012). Similarly, cognitive and emotional processes may help explain the association between psychophysiological risk and engagement in relationally aggressive conduct. In fact, several theoretical perspectives suggest that physiological stress arousal is related to aggression because it reflects emotional experiences of anger, fear, or anxiety (see Murray-Close, 2013a). Physiological arousal may also lead to aggression by interfering with accurate cognitive processing of threat or provocation. In fact, Williams, Lochman, Phillips, and Barry (2003) found that increases in heart rate following threat were associated with increases in HAB. These results are consistent with the suggestion that exaggerated negative emotional responses to stress lead to socialcognitive impairments that promote involvement in aggressive behavior (Frick & Morris, 2004). Biological factors, including genetic and physiological risk, may also lead to increases in relationally aggressive conduct because they alter children's social environments. In fact, theorists have suggested several processes whereby children's genetic makeup affects the social environments that they encounter (Brendgen, 2012). For instance, genetically influenced traits may lead children to actively select specific environments (i.e., active gene-environment correlations) and/or to elicit distinct reactions from others (i.e., evocative gene-environment correlations) that in turn increase their risk for involvement in relationally aggressive conduct (Brendgen, 2012). As an example, a child who has genetic risk for emotion dysregulation may

be more likely to choose aggressive friends and to elicit rejection from peers; these social experiences, in turn, have been found to increase risk for relationally aggressive conduct (see the “Risk Factors: Social Processes” section). Similarly, a tendency to experience exaggerated physiological arousal during peer conflict situations may make youth easy targets of peer victimization (Murray-Close, 2013b), a social stressor that has been found to increase aggressive responses (see the “Risk Factors: Social Processes” section). Mechanisms of influence across multiple levels of analysis are likely bidirectional in nature; thus, alterations in genetic and physiological processes may also serve as mediators of the association between social experiences and relationally aggressive behavior. For instance, exposure to peer victimization may modify gene expression through epigenetic processes, thus increasing risk for social behaviors such as aggression (Brendgen, 2012). Similarly, an emerging body of research has documented that victimization by peers is associated with alterations in stress physiology (see Murray-Close, 2013a, for review); these findings highlight the possibility that negative peer treatment may lead to physiological profiles that in turn promote relationally aggressive conduct. Social-cognitive risk factors for aggression, such as HAB, may also help explain why social experiences such as peer victimization and media exposure are related to elevated levels of aggressive conduct (e.g., Gentile, Coyne, et al., 2011; Hoglund & Leadbeater, 2007; Ostrov et al., 2011). Mediation models across levels of analysis may further help elucidate the mechanisms that link relational aggression with negative developmental outcomes, such as peer victimization (e.g., Ostrov, 2008; Ostrov & Godleski, 2013) and psychopathology (Kamper & Ostrov, 2013). Although preliminary, these findings highlight the critical need for multilevel, transactional research examining the processes linking risk factors to involvement in relationally aggressive conduct. Recent statistical advances, including developmental cascades models, allow for longitudinal tests of how functioning in one area can spill over into other domains of functioning (Masten et al., 2005). Developmental cascades approaches that span multilevel domains of functioning (see Cox, Mills-Koonce, Propper, & Gariépy, 2010) may be especially useful in enhancing our understanding of the longitudinal processes involved in the development of relational aggression. Moderation Models Multilevel perspectives regarding the development of relational aggression also highlight the possibility that risk factors may interact across levels of analysis in the prediction of relationally aggressive conduct. In fact, a large body of research regarding aggression has emphasized a diathesis–stress model of aggression, whereby an underlying trait or vulnerability is activated through exposure to environmental stressors. Although most work in this area has focused on physical forms of aggression, several recent findings highlight the potential role of diathesis–stress processes in the development of relational forms of aggression. For instance, Mathieson et al. (2011) reported that girls with cognitive (i.e., HAB) and emotional (i.e., emotional reactivity to provocation) vulnerability for relational aggression were most likely to engage in such conduct when they were exposed to social risk (i.e., relational victimization by peers). Diathesis–stress models have also been employed in the

study of physiological risk for relational aggression. For instance, Murray-Close and Rellini (2012) found that physiological risk was associated with relational aggression only among women with a history of sexual abuse, a contextual factor that has been found to predict relational forms of aggression in previous research (Cullerton-Sen et al., 2008). Similarly, Murray-Close (2011) found that physiological risk was most strongly associated with romantic relational aggression among women who also exhibited cognitive (i.e., HAB) and social (i.e., relational victimization) vulnerabilities for such conduct. It will be important for future research to examine the diathesis–stress model in the context of genetic risk factors for relational aggression. In fact, G×E interactions highlight how genetic vulnerabilities may result in aggressive behavior only when they interact with environmental stressors, such as maltreatment by parents or peers (Brendgen, 2012). To date, little research has explored gene-environment interactions in the development of relational aggression. In preliminary research in this area, Brendgen et al. (2008) found that genetic risk interacted with friends' aggressive behavior in the prediction of physical, but not social, forms of aggression. The authors suggest that children may be likely to emulate social aggression from friends, regardless of genetic vulnerability, given the relatively low costs (e.g., low likelihood of adult intervention) and potential benefits (e.g., high social status) associated with social aggression (Brendgen et al., 2008). However, it will be important for future research to further explore diathesis–stress models regarding genetic vulnerabilities for relational aggression. Such work may benefit from adopting molecular genetic designs and investigating candidate genes implicated in the development of physical aggression and antisocial behaviors (see Brendgen, 2012). It will also be important for this work to focus on the unique social stressors that may activate genetic risk for relational aggression; for instance, environmental pathogens that are uniquely associated with the development of relational aggression (e.g., parental psychological control; see the “Risk Factors: Social Processes” section) may be particularly important to explore in studies of G×E interactions. In contrast to diathesis–stress models of relational aggression, some researchers have proposed that biological vulnerabilities may be most likely to translate into aggressive or antisocial behavior in the context of benign environmental contexts (e.g., high SES). More specifically, the social push hypothesis suggests that adverse environments may overwhelm biological risk for aggression; in these contexts, social risk factors may be the strongest predictors of such conduct. In contrast, genetic and physiological differences may exert a stronger effect on aggressive behavior in social contexts that do not promote aggression. This perspective is similar to the discussion of suppression effects in genetic studies, in which some environmental factors reduce the role of genetic influences on behavior (Brendgen, 2012). Although several studies have provided support for the social push perspective in the development of physical aggression and antisocial behavior (see Raine, 2002, for a review), there is less support for this model in the development of relational forms of aggression. One exception is a study documenting that blunted morning cortisol and flattened change in cortisol across the day were more strongly associated with relational aggression in nonmaltreated, as compared to maltreated, youth (Murray-Close et al., 2008). It is important to note that both maltreated and nonmaltreated participants in this study experienced significant social risk (e.g.,

low income), highlighting the possibility that severe environmental contexts (e.g., abuse) may be more likely than moderate stressors (e.g., peer victimization) to overwhelm biological factors in the development of relational aggression. In recent years, researchers have proposed that individuals differ in their susceptibility to environmental influence, such that those who exhibit heightened physiological reactivity to stress, are more malleable or plastic, for better or worse (Boyce & Ellis, 2005). This theoretical approach is an outgrowth of the diathesis–stress model but extends the idea to argue that the very same individuals who are most negatively affected by adverse environmental input will also benefit most from supportive contexts (Ellis et al., 2011). Thus, individual differences in biological processes that have been proposed as risk factors for aggression (e.g., high physiological stress reactivity) may in fact reflect the organism's ability both to prepare for challenges in the face of adversity as well as to benefit from the resources available in supportive contexts (Boyce & Ellis, 2005). This hypothesis suggests that it will be important for studies to include assessments of positive social contexts (e.g., relational inclusion) to examine whether biological factors that serve as risk factors for relational aggression in adverse contexts portend beneficial outcomes in supportive contexts.

Equifinality and Multifinality Developmental psychopathologists have also highlighted the importance of examining pathways to maladaptive behavior patterns (Cicchetti & Rogosch, 1996; Egeland, Pianta, & Ogawa, 1996; Sroufe, 1997). The concept of equifinality suggests that a particular outcome may be achieved via a number of distinct pathways (Cicchetti & Rogosch, 1996). This notion highlights the possibility that there may be multiple trajectories, each with a different set of risk factors, to relational aggression. As discussed, researchers have documented two distinct physiological profiles associated with relational aggression; on one hand, some evidence indicates that low-stress reactivity, hypothesized to reflect temperamental fearlessness, predicts heightened involvement in relationally aggressive conduct (e.g., Sijtsema et al., 2011). On the other hand, findings from other studies suggest that relational aggression is related to physiological overarousal in response to negative experiences with peers, perhaps because this pattern reflects angry reactions to peer conflict (Murray-Close & Crick, 2007). The concept of equifinality underscores the possibility that both suggestions may be correct; in effect, blunted and exaggerated physiological reactivity in response to peer stress may promote phenotypically similar displays of relational aggression via distinct trajectories (see Frick & Morris, 2004, for a discussion of these pathways in the development of physical aggression). An important task for researchers in this area, then, is to identify the underlying processes as well as the moderating factors that channel youth into each pathway. For example, heightened physiological reactivity may promote relational aggression when it occurs in individuals who are frequently exposed to contextual factors that elicit stress reactions (e.g., victimization) and who interpret these stressful events as intentional and hostile (e.g., exhibit HAB; MurrayClose, 2011). In contrast, blunted physiological reactivity may promote relational aggression among individuals with relatively low levels of HAB (Murray-Close, 2011). The related concept of multifinality, which suggests that a single etiological factor can result

in multiple developmental outcomes (Cicchetti & Rogosch, 1996), may also inform the study of relational aggression in several ways. First, researchers frequently argue that relational and physical forms of aggression are both manifestations of externalizing behaviors (Crick & ZahnWaxler, 2003; Tackett et al., 2009). From this perspective, similar underlying vulnerabilities or traits may promote both forms of aggression, with specific manifestations depending on factors such as the gender, age, or social goals of the aggressor (Burt et al., 2012; Crick & Zahn-Waxler, 2003; Lagerspetz & Björkqvist, 1994). In fact, some studies have demonstrated overlapping genetic (Brendgen et al., 2005), social (e.g., maltreatment; Cullerton-Sen et al., 2008), and personality (Burt et al., 2012) risk factors for physical and relational aggression. In future work in this area, it will be particularly important to identify moderating factors (e.g., gender; see Cullerton-Sen et al., 2008, for an example) that can help explain the circumstances or contexts in which a shared risk factor results in the manifestation of relationally, rather than physically, aggressive behaviors. Additionally, as some studies have found unique risk factors for physical versus relational forms of aggression (e.g., Godleski et al., 2010; Sijtsema et al., 2011), theoretical and empirical work must also consider distinct developmental pathways toward such conduct. In effect, it may not always be appropriate to use theory and past research on physical aggression to justify predictions regarding relational aggression. The concept of multifinality also highlights the possibility that behavioral displays of relational aggression may be related to distinct sets of developmental outcomes. As reviewed earlier, a large body of research indicates that youth who engage in relational aggression are at risk for social, psychological, and behavioral maladjustment. However, there has been considerable debate regarding whether relational aggression may at times serve adaptive functions, such as allowing youth to establish dominant and high-status positions within the peer group (e.g., Hawley et al., 2007). In fact, an emerging body of research indicates that some relationally aggressive youth enjoy relatively high levels of social power or prominence with peers, depending on factors such as their gender, age, and prosocial behavior (Cillessen & Mayeux, 2004; Hawley, 2003a; Hawley et al., 2007). These findings are consistent with the concept of multifinality and suggest that the meaning of relationally aggressive conduct may be different for youth depending on where they fall on a constellation of other traits. Moreover, findings highlight the potential utility of person-centered approaches in identifying profiles of relationally aggressive youth who experience negative versus positive developmental outcomes (see Hawley et al., 2007, 2008). An important extension of this work will be to assess the relative dispersion of outcomes (i.e., multifinality) associated with relational aggression as compared to other facets of externalizing pathology or antisocial behavior (see Egeland et al., 1996, for a comparison of dispersion for early internalizing and externalizing pathology).

Future Directions Antisocial behavior is one of the most well-researched topics in developmental psychopathology, and the inclusion of relational aggression in this work has dramatically altered our understanding of aggressive behavior problems in youth. Since Crick and

Grotpeter's (1995) seminal publication on the topic of relational aggression two decades ago, several important directions for future research have emerged.

Understanding How Relational Aggression Relates to Antisocial Behaviors Critical questions remain regarding the overlap between forms of aggressive behavior. For instance, although social, indirect, and relational forms of aggression are conceptually similar, we do not believe that these terms are interchangeable (see the “Defining Relational Aggression” section). There may be theoretically meaningful differences in the etiology and outcomes associated with each subtype of behavior; relational aggression, for example, may be particularly likely to occur during early childhood and may be more strongly tied to close, intimate friendships and other close relationship systems. Researchers also should explore relational bullying, defined as repeated perpetration of relationally aggressive acts in the context of a relationship characterized by a power imbalance (e.g., the target is physically smaller or less popular than the aggressor; Gladden, Vivolo-Kantor, Hamburger, & Lumpkin, 2014; Terranova, Morris, & Boxer, 2008). Researchers have yet to evaluate the extent to which relational aggression and relational bullying are theoretically and empirically distinct. It will be particularly important to isolate effects that are specific to relational bullying versus more general acts of relational aggression. For example, much work on relational aggression in the context of close friendships likely does not address relational bullying. In contexts that may include both relational aggression and relational bullying (e.g., classroom peer relationships), it is unclear whether relational bullying is associated with more problematic outcomes than relational aggression. Moreover, initial studies suggest that relational aggression exhibits significant overlap with electronic forms of aggression (Werner, Bumpus, & Rock, 2010), with some longitudinal evidence indicating that relational aggression precedes electronic aggression (e.g., Hemphill et al., 2012; Werner et al., 2010). A critical question for future research is whether relational and electronic aggression are distinct or whether they are analogous behaviors manifested in different formats (i.e., face to face versus using electronic media; see Wright & Li, 2013). Future research must also evaluate the degree to which relational aggression is best conceptualized and assessed as a distinct behavior versus a manifestation of antisocial behavior and externalizing pathology (e.g., Burt et al., 2012; Crick & Zahn-Waxler, 2003; Tackett, Daoud, De Bolle, & Burt, 2013). Researchers often integrate measures of relational aggression into instruments assessing a broader range of aggressive or antisocial conduct, a practice that is supported by evidence indicating that relational aggression: (1) is highly correlated with other aggressive behaviors (e.g., physical aggression; Card et al., 2008); (2) significantly loads on a single antisocial behavior factor with physical aggression and rulebreaking behaviors (Tackett et al., 2013); and (3) exhibits similar correlations with personality factors as other facets of antisocial conduct (Tackett et al., 2013). Nevertheless, as reviewed in this chapter, some evidence suggests that relational and physical aggression have unique developmental precursors and outcomes. These findings underscore the critical need for empirical tests of whether the developmental trajectories associated with relational aggression

are similar to (or distinct from) trajectories of other indices of antisocial conduct, such as physical aggression. We believe that a top priority for the next decade of research on relational aggression should be to improve the conceptual clarity in the field regarding how relational aggression relates to various facets of antisocial behavior. To this end, we have several suggestions for future work in this area. First, it is especially important for future researchers to adopt the terminology that was originally intended for the measures used (e.g., relational aggression versus relational bullying versus indirect aggression). Second, researchers must examine the empirical overlap across measures (and across items within measures) to test the extent to which these constructs are similar versus distinct. One practical implication of this recommendation is that researchers must include multiple items assessing each subtype of behavior rather than integrating single items of relational aggression into broader assessments of antisocial conduct. New studies should be developed with the aim of testing the shared and distinct causes and correlates of relational aggression versus other forms of aggression and antisocial behavior; incorporating multiple facets of antisocial behavior in these studies will be essential to understanding the processes of equifinality and multifinality (see Egeland et al., 1996). Moreover, researchers should evaluate the extent to which interventions exhibit specificity in reducing different behaviors (e.g., physical aggression, relational aggression, relational bullying, and electronic aggression). This work will have critical implications for applied efforts to reduce aggressive and antisocial conduct. For instance, given emerging research highlighting the unique risk factors for relational aggression, it may be necessary to develop additional interventions that specifically target relational aggression because methods and approaches developed for physical aggression or other antisocial behaviors may not be efficacious.

Moving Beyond “Mean Girls” In early research, relational aggression was often conceptualized as a behavior primarily enacted by child and adolescent girls (see Crick & Grotpeter, 1995). This has led to the popular notion that relational aggression is a problem of mean girls. Consistent with this perspective, the inclusion of relational forms of aggression is essential to research on the aggressive behavior patterns of girls, both because physical aggression is uncommon among girls and because the overlap between these two forms of aggression appears to be lower for girls than for boys. Thus, when girls are aggressive, they tend to engage in relational, but not physical, forms of aggression. A failure to include measures of relational aggression will lead researchers to overlook the vast majority of aggressive females. At the same time, mounting evidence has highlighted the need for research to expand beyond these circumscribed gender and age boundaries. For instance, relational aggression is not only perpetrated by females; evidence indicates that males engage in these behaviors with some frequency. Moreover, the vast majority of research on relational aggression has concentrated on these behaviors during early childhood through adolescence. This narrow focus has likely emerged from the persistent view that relational aggression is problem of adolescence, a perspective that has been allowed to thrive in the absence of good estimates of the incidence of

relational aggression across developmental periods. Nonetheless, as relational aggression is theorized to promote social power and resource control in groups (e.g., Hawley, 2003a), such conduct may emerge with some frequency in adult contexts, such as the workplace. For instance, relational aggression may be used to climb the corporate ladder or to denigrate career competitors. Despite these possibilities, we know little about the use of relational aggression in dyadic (e.g., close friendships and romantic relationships) and group-based relational contexts (e.g., the workplace) during adulthood. As researchers expand the study of relational aggression beyond mean girls, we have several recommendations regarding the most critical research priorities. First, we need epidemiological studies regarding the prevalence of relational aggression in males and females across developmental periods. Second, researchers should examine the use of relational aggression in adulthood (particularly beyond the college years); such work has the potential to provide important insights into adult adjustment and inform interventions aimed at improving group dynamics (e.g., workplace productivity) and functioning in close relationships (e.g., marital harmony). Third, researchers should conduct longitudinal studies of the long-term impact of continuity in relationally aggressive behavior. Studies that assess trajectories of relational aggression (see Cleverley et al., 2012, for an excellent example) have the potential to identify why some individuals desist from such conduct as well as inform the debate regarding the developmental outcomes of relational aggression. In fact, some of the costs and benefits of relational aggression may emerge during the adult years (see Sandstrom & Cillessen, 2010). Finally, it is critical for researchers to continue to explore the role of gender in the correlates of relational aggression as the meaning and consequences of such conduct may be gender-linked (see Ostrov & Godleski, 2010).

Developing New Methods and Research Designs to Study Relational Aggression We believe that an essential direction for the next generation of relational aggression research is the inclusion of novel measures and research designs. For instance, few approaches beyond behavioral observations allow researchers to capture relationally aggressive behaviors in vivo. Techniques that allow researchers to study relational aggression while it is emerging, including potential factors that precipitate these behaviors as well as their immediate outcomes, would provide researchers with a more fine-grained understanding of the transactional processes underlying such conduct. It is particularly important for researchers to move beyond trait-based conceptualizations of relationally aggressive behavior and associated outcomes (e.g., reactively relationally aggressive youth tend to be victimized) to examine the dynamics of specific interactions using the time scale in which these processes are most likely to occur (e.g., a peer provokes the child, the child responds with reactive relational aggression, the peer then ignores the child, etc.; see Orobio de Castro, Thomaes, & Reijntjes, 2015). Such approaches would provide a rigorous test of the transactional processes theorized to underlie relational aggression and would allow researchers to link long-term developmental change with short-term exchanges (see van Geert & Steenbeek, 2005).

It is also critical for researchers to adopt methodological approaches that can test the causal pathways hypothesized to underlie relational aggression (see Orobio de Castro et al., 2015, for a discussion of integrating experimental designs into peer relationships research). To date, the vast majority of work in this area is correlational and cross-sectional in nature; the incorporation of experimental designs promises to further clarify the mechanisms associated with relational aggression. Hypothesized developmental precursors of relational aggression that can be manipulated, such as exposure to aggressive media, should be examined in the context of experimental research designs (e.g., Coyne et al., 2008). Evaluation of the effects of interventions may also provide insight into the causes of relational aggression, particularly in contexts where ethical constraints prohibit random assignment to experimental conditions (e.g., child maltreatment). Finally, factors that cannot be directly manipulated (e.g., culture, genes, gender) can be included in experimental designs as moderators. Long-term longitudinal designs that incorporate rigorous controls of earlier functioning (e.g., developmental cascades; Masten et al., 2005) should also be employed to better clarify directional associations between relational aggression and putative risk factors and outcomes. We believe that researchers should increasingly incorporate person-centered analyses (e.g., latent profile analysis, latent class growth analysis) to identify heterogeneity among relationally aggressive youth. For instance, emerging evidence indicates that there may be distinct subgroups of relational aggressors: Some perpetrators appear to enjoy social power and status (e.g., popularity) whereas others suffer poor social outcomes (e.g., rejection). The inclusion of person-centered analyses may help clarify distinct pathways to relational aggression (equifinality) and the differing developmental outcomes related to such conduct (multifinalty; see Cicchetti & Rogosch, 1996). Further, this research could be integrated into intervention work: Interventions for popular relational aggressors may need to identify ways to replace the social status benefits of aggression in a manner that facilitates youth buy-in (Leff, Waasdorp & Crick, 2010). Interventions with low-status youth, in contrast, may be most efficacious when they promote social skill development.

Culture and Context Extant research has highlighted the critical role of context in the development of relational aggression. It is essential for future researchers to broaden this focus to encompass a wider range of relational and cultural contexts, including relationship types (e.g., friendships, siblings), countries, SES, urban/rural environments, school contexts, and religious and familial value systems. In fact, although there are an impressive number of studies of relational aggression in countries other than the United States, explorations of how these behaviors differ across multiple countries or across diverse populations within a single country are still relatively rare. We believe that a top priority for this work should be the adoption of multilevel perspectives that investigate interactions and bidirectional associations across various levels of analysis (i.e., from genes/cells to communities). Importantly, studies of contextual effects should move beyond a singular focus on differences in the frequency of relational aggression to consider differences in the meaning of such conduct across these contexts (Kawabata et al., 2010). For example, it is crucial to investigate whether pathways to relational aggression and

associated developmental outcomes vary as a function of distinct contexts (see Kawabata et al., 2010, for an example). It is also important for researchers to investigate the mechanisms hypothesized to link contexts to relationally aggressive behavior (i.e., mediation models). As an example, some work has documented SES differences in the use of relational (or indirect) aggression, although the direction of effects has varied across studies (e.g., Bonica et al., 2003; Vaillancourt et al., 2007). In an effort to understand the processes that lead to these differences, it is necessary to integrate factors that may be related to SES and that have been implicated in the development of relational aggression (e.g., parenting behaviors, family stress and conflict; see the “Risk Factors: Social Processes” section). These approaches may also inform our understanding of within-group differences in relational aggression; for instance, there is likely substantial heterogeneity in parenting behaviors within SES backgrounds, and these differences may in turn have important implications for children's involvement in relational aggression. In sum, we believe that there are several critical future directions that have the potential to advance the study of relational aggression in new and exciting directions. As we embark on the next generation of relational aggression research, it will be crucial to engage in significant theory building and to ground new research studies in theory. Additionally, given the clear need for direct replication in psychological research (see Pashler & Harris, 2012), new approaches to the study of relational aggression should be augmented with studies developed to replicate early relational aggression findings.

Conclusion Research on relational aggression has burgeoned in recent years, and this work has significantly altered the ways in which aggressive and antisocial behaviors are conceptualized both in research and in practice. Theoretically, the construct of relational aggression has forced developmental psychopathologists to reconsider the long-standing assumption that girls do not exhibit significant behavioral problems prior to adolescence. Moreover, this work has challenged parents, teachers, and clinicians to seriously consider the developmental ramifications of relational aggression. Despite these significant advances, critical questions remain. For instance, why are some relationally aggressive children highly rejected, whereas others hold coveted positions in the peer group? When is relational aggression a normative behavior, and when does it serve as an index of risk for poor developmental outcomes? How can parents, teachers, and clinicians intervene with perpetrators and victims of relational aggression? We believe that the theoretical, methodological, and data-analytic advances in the field of developmental psychopathology hold great promise in helping researchers address these and other significant questions regarding the etiology and developmental outcomes of relational aggression.

References Adams, R. E., Bartlett, N. H., & Bukowski, W. M. (2010). Peer victimization and social

dominance as intervening variables of the link between peer liking and relational aggression. Journal of Early Adolescence, 30, 102–121. doi: 10.1177/0272431609342985 Agoston, A. M. & Rudolph, K. D. (2013). Pathways from depressive symptoms to low social status. Journal of Abnormal Child Psychology, 41(2), 295–308. doi: 10.1007/s10802–012– 9675-y Allport, G. W., Bruner, J. S., & Jandorf, E. M. (1941). Personality under social catastrophe: Ninety life-histories of the Nazi revolution. Charter & Personality: A Quarterly for Psychodiagnostic & Allied Studies, 10, 1–22. Anderson, C. A., & Bushman, B. J. (2001). Effects of violent video games on aggressive behavior, aggressive cognition, aggressive affect, physiological arousal, and prosocial behavior: A meta-analytic review of the scientific literature. Psychological Science, 12(5), 353–359. doi: 10.1111/1467–9280.00366 Anderson, C. A., & Bushman, B. J. (2002). Human aggression. Annual Review of Psychology, 53, 27–51. Anderson, C. A., Shibuya, A., Ihori, N., Swing, E. L., Bushman, B. J., Sakamoto, A.,…Saleem, M. (2010). Violent video game effects on aggression, empathy, and prosocial behavior in Eastern and Western countries: A meta-analytic review. Psychological Bulletin, 136(2), 151– 173. doi: 10.1037/a0018251 Andreou, E. (2006). Social preference, perceived popularity and social intelligence: Relations to overt and relational aggression. School Psychology International, 27(3), 339–351. doi: 10.1177/0143034306067286 Archer, J. (2006). Testosterone and human aggression: An evaluation of the challenge hypothesis. Neuroscience & Biobehavioral Reviews, 30(3), 319–345. doi: 10.1016/j.neubiorev.2004.12.007 Archer, J., & Coyne, S. M. (2005). An integrated review of indirect, relational, and social aggression. Personality and Social Psychology Review, 9(3), 212–230. doi: 10.1207/s15327957pspr0903_2 Archer, J., Graham-Kevan, N., & Davies, M. (2005). Testosterone and aggression: A reanalysis of Book, Starzyk, and Quinsey's (2001) study. Aggression and Violent Behavior, 10(2), 241–261. doi: 10.1016/j.avb.2004.01.001 Arsenault, D. J., & Foster, S. L. (2012). Attentional processes in children's overt and relational aggression. Merrill-Palmer Quarterly, 58(3), 409–436. doi: 10.1353/mpq.2012.0015 Bagner, D. M., Storch, E. A., & Preston, A. S. (2007). Romantic relational aggression: What about gender? Journal of Family Violence, 22, 19–24. doi: 10.1007/s10896–006–9055-x Bailey, C. A., & Ostrov, J. M. (2008). Differentiating forms and functions of aggression in

emerging adults: Associations with hostile attribution biases and normative beliefs. Journal of Youth & Adolescence, 37(6), 713–722. doi: 10.1007/s10964–007–9211–5 Baird, A. A., Silver, S. H., & Veague, H. B. (2010). Cognitive control reduces sensitivity to relational aggression among adolescent girls. Social Neuroscience, 5(5–6), 519–532. doi: 10.1080/17470911003747386 Bandura, A. (1973). Aggression: A social learning analysis. Englewood Cliffs, NJ: Prentice Hall. Banny, A. M., Heilbron, N., Ames, A., & Prinstein, M. J. (2011). Relational benefits of relational aggression: Adaptive and maladaptive associations with adolescent friendship quality. Developmental Psychology, 47(4), 1153–1166. doi: 10.1037/a0022546 Banny, A. M., Tseng, W.-L., Murray-Close, D., Pitula, C., & Crick, N. R. (2014). Borderline personality features as a predictor of forms and functions of aggression during middle childhood: Examining the roles of gender and physiological reactivity. Development and Psychopathology, 26, 789–804. doi: 10.1017/S095457941400039X Barry, C. T., Grafeman, S. J., Adler, K. K., & Pickard, J. D. (2007). The relations among narcissism, self-esteem, and delinquency in a sample of at-risk adolescents. Journal of Adolescence, 30(6), 933–942. doi: 10.1016/j.adolescence.2006.12.003 Barry, C. T., Pickard, J. D., & Ansel, L. L. (2009). The associations of adolescent invulnerability and narcissism with problem behaviors. Personality and Individual Differences, 47(6), 577–582. doi: 10.1016/j.paid.2009.05.022 Batanova, M. D., & Loukas, A. (2011). Social anxiety and aggression in early adolescents: Examining the moderating roles of empathic concern and perspective taking. Journal of Youth and Adolescence, 40(11), 1534–1543. doi: 10.1007/s10964–011–9634-x Becker, S. P., Luebbe, A. M., Stoppelbein, L., Greening, L., & Fite, P. J. (2012). Aggression among children with ADHD, anxiety, or co-occurring symptoms: Competing exacerbation and attenuation hypotheses. Journal of Abnormal Child Psychology, 40(4), 527–542. doi: 10.1007/s10802–011–9590–7 Belden, A. C., Gaffrey, M. S., & Luby, J. L. (2012). Relational aggression in children with preschool-onset psychiatric disorders. Journal of the American Academy of Child & Adolescent Psychiatry, 51(9), 889–901. doi: 10.1016/j.jaac.2012.06.018 Belgrave, F. Z., Reed, M. C., Plybon, L. E., Butler, D. S., Allison, K. W., & Davis, T. (2004). An evaluation of Sisters of Nia: A cultural program for African American girls. Journal of Black Psychology, 30(3), 329–343. doi: 10.1177/0095798404266063 Berkowitz, L. (1962). Aggression: A social psychological analysis. New York, NY: McGrawHill.

Berndt, T. J. (1996). Transitions in friendship and friends' influence. In J. A. Graber, J. Brooks-Gunn, & A. C. Petersen (Eds.), Transitions through adolescence: Interpersonal domains and context (pp. 57–84). Mahwah, NJ: Erlbaum. Bierman, K. L. (2004). Peer rejection: Developmental processes and intervention strategies. New York, NY: Guilford Press. Björkqvist, K. (1994). Sex differences in physical, verbal, and indirect aggression: A review of recent research. Sex Roles, 30(3–4), 177–188. doi: 10.1007/BF01420988 Björkqvist, K., Lagerspetz, K. M., & Kaukiainen, A. (1992). Do girls manipulate and boys fight? Developmental trends in regard to direct and indirect aggression. Aggressive Behavior, 18(2), 117–127. doi: 10.1002/1098–2337(1992)18:2 Blake, J. J., Kim, E. S., & Lease, M. A. (2011). Exploring the incremental validity of nonverbal social aggression: The utility of peer nominations. Merrill-Palmer Quarterly, 57(3), 293–318. doi: 10.1353/mpq.2011.0015 Bonica, C., Arnold, D. H., Fisher, P. H., Zeljo, A., & Yershova, K. (2003). Relational aggression, relational victimization, and language development in preschoolers. Social Development, 12(4), 551–562. doi: 10.1111/1467–9507.00248 Bowker, J. C., Ostrov, J. M., & Raja, R. (2012). Relational and overt aggression in urban India: Associations with peer relations and best friends' aggression. International Journal of Behavioral Development, 36(2), 107–116. doi: 10.1177/0165025411426019 Boyce, W. T., & Ellis, B. J. (2005). Biological sensitivity to context: I. An evolutionarydevelopmental theory of the origins and functions of stress reactivity. Development and Psychopathology, 17(2), 271–301. doi: 10.1017/S0954579405050145 Brendgen, M. (2012). Genetics and peer relations: A review. Journal of Research on Adolescence, 22(3), 419–437. doi: 10.1111/j.1532–7795.2012.00798.x Brendgen, M., Boivin, M., Vitaro, F., Bukowski, W. M., Dionne, G., Tremblay, R. E., & Pérusse, D. (2008). Linkages between children's and their friends' social and physical aggression: Evidence for a gene-environment interaction? Child Development, 79, 13–29. doi: 10.1111/j.1467–8624.2007.01108.x Brendgen, M., Dionne, G., Girard, A., Boivin, M., Vitaro, F., & Pérusse, D. (2005). Examining genetic and environmental effects on social aggression: A study of 6-year-old twins. Child Development, 76(4), 930–946. doi: 10.1111/j.1467–8624.2005.00887.x Brown, S. A., Arnold, D. H., Dobbs, J., & Doctoroff, G. L. (2007). Parenting predictors of relational aggression among Puerto Rican and European American school-age children. Early Childhood Research Quarterly, 22, 147–159. doi: 10.1016/j.ecresq.2006.11.002 Browne, D. T., Odueyungbo, A., Thabane, L., Byrne, C., & Smart, L. A. (2010). Parenting-by-

gender interactions in child psychopathology: Attempting to address inconsistencies with a Canadian national database. Child and Adolescent Psychiatry and Mental Health, 4, 5–17. doi: 10.1186/1753–2000–4–5 Bukowski, W. M., Schwartzman, A., Santo, J., Bagwell, C., & Adams, R. (2009). Reactivity and distortions in the self: Narcissism, types of aggression, and the functioning of the hypothalamic-pituitary-adrenal axis during early adolescence. Development and Psychopathology, 21(4), 1249–1262. doi: 10.1017/S0954579409990149 Burnette, M. L., & Reppucci, N. D. (2009). Childhood abuse and aggression in girls: The contribution of borderline personality disorder. Development and Psychopathology, 21, 309– 317. doi: 10.1017/S0954579409000170 Burr, J. E., Ostrov, J. M., Jansen, E. A., Cullerton-Sen, C., & Crick, N. R. (2005). Relational aggression and friendship during early childhood: “I won't be your friend!” Early Education and Development, 16(2), 161–183. doi: 10.1207/s15566935eed1602_4 Burt, S. A., Donnellan, M. B., & Tackett, J. L. (2012). Should social aggression be considered “antisocial”? Journal of Psychopathology and Behavioral Assessment, 34(2), 153–163. doi: 10.1007/s10862–011–9267–0. Buss, A. H. (1961). The Psychology of Aggression. New York, NY: Wiley. Butovskaya, M. L., Timentschik, V. M., & Burkova, V. N. (2007). Aggression, conflict resolution, popularity, and attitude to school in Russian adolescents. Aggressive Behavior, 33(2), 170–183. doi: 10.1002/ab.20197 Cairns, R. B., Cairns, B. D., Neckerman, H. J., Ferguson, L. L., & Gariépy, J. (1989). Growth and aggression: 1. Childhood to early adolescence. Developmental Psychology, 25(2), 320– 330. doi: 10.1037/0012–1649.25.2.320 Calkins, S. D., & Keane, S. P. (2004). Cardiac vagal regulation across the preschool period: Stability, continuity, and implications for childhood adjustment. Developmental Psychobiology, 45(3), 101–112. doi: 10.1002/dev.20020 Campione-Barr, N., Lindell, A. K., Greer, K. B., & Rose, A. J. (2014). Relational aggression and psychological control in the sibling relationship: Mediators of the association between maternal psychological control and adolescents' emotional adjustment. Development and Psychopathology, 26, 749–758. doi: 10.1017/S0954579414000364 Card, N. A., & Little, T. D. (2006). Proactive and reactive aggression in childhood and adolescence: A meta-analysis of differential relations with psychosocial adjustment. International Journal of Behavioral Development, 30(5), 466–480. doi: 10.1177/0165025406071904 Card, N. A., Stucky, B. D., Sawalani, G. M., & Little, T. D. (2008). Direct and indirect aggression during childhood and adolescence: A meta-analytic review of gender differences,

intercorrelations, and relations to maladjustment. Child Development, 79, 1185–1229. doi: 10.1111/j.1467–8624.2008.01184.x Carnagey, N. L., & Anderson, C. A. (2003). Theory in the study of media violence: The general aggression model. In D. A. Gentile (Ed.). Media violence and children (pp. 87–105). Westport, CT: Praeger. Carroll, J. S., Nelson, D. A., Yorgason, J. B., Harper, J. M., Ashton, R. H., & Jensen, A. C. (2010). Relational aggression in marriage. Aggressive Behavior, 36(5), 315–329. doi: 10.1002/ab.20349 Casas, J. F., & Albers, S. M. (2014). Relational and electronic aggression and victimization: An examination across development and contexts. Poster presented at the biennial conference of the Society for Research on Adolescence, Austin, TX. Casas, J. F., Weigel, S. M., Crick, N. R., Ostrov, J. M., Woods, K. E., Yeh, E. A. J., & Huddleston-Casas, C. A. (2006). Early parenting and children's relational and physical aggression in the preschool and home contexts. Journal of Applied Developmental Psychology, 27(3), 209–227. doi: 10.1016/j.appdev.2006.02.003 Cicchetti, D. (1993). Developmental psychopathology: Reactions, reflections, projections. Developmental Review, 13, 471–502. doi: 10.1006/drev.1993.1021 Cicchetti, D., & Crick, N. R. (2009). Editorial: Precursors and diverse pathways to personality disorder in children and adolescents. Development and Psychopathology, 21, 683–685. doi: 10.1017/S0954579409000388 Cicchetti, D., & Rogosch, F. A. (1996). Equifinality and multifinality in developmental psychopathology. Development and Psychopathology, 8(4), 597–600. doi: 10.1017/S0954579400007318 Cillessen, A. H. N., Jiang, X. L., West, T. V., & Laszkowski, D. K. (2005). Predictors of dyadic friendship quality in adolescence. International Journal of Behavioral Development, 29(2), 165–172. doi: 10.1080/01650250444000360 Cillessen, A. H. N., & Mayeux, L. (2004). From censure to reinforcement: Developmental changes in the association between aggression and social status. Child Development, 75, 147– 163. doi: 10.1111/j.1467–8624.2004.00660.x Cleverley, K., Szatmari, P., Vaillancourt, T., Boyle, M., & Lipman, E. (2012). Developmental trajectories of physical and indirect aggression from late childhood to adolescence: Sex differences and outcomes in emerging adulthood. Journal of the American Academy of Child & Adolescent Psychiatry, 51(10), 1037–1051. doi: 10.1016/j.jaac.2012.07.010 Côté, S. M., Vaillancourt, T., Barker, E. D., Nagin, D., & Tremblay, R. E. (2007). The joint development of physical and indirect aggression: Predictors of continuity and change during childhood. Development and Psychopathology, 19, 37–55. doi:

10.1017/S0954579407070034 Cox, M. J., Mills-Koonce, R., Propper, C., & Gariépy, J.-L. (2010). Systems theory and cascades in developmental psychopathology. Development and Psychopathology, 22(3), 497– 506. doi: 10.1017/S0954579410000234 Coyne, S. M., Archer, J., & Eslea, M. (2004). Cruel intentions on television and in real life: Can viewing indirect aggression increase viewers' subsequent indirect aggression? Journal of Experimental Child Psychology, 88(3), 234–253. doi: 10.1016/j.jecp.2004.03.001 Coyne, S. M., Archer, J., & Eslea, M. (2006). “We're not friends anymore! unless…”: The frequency and harmfulness of indirect, relational, and social aggression. Aggressive Behavior, 32(4), 294–307. doi: 10.1002/ab.20126 Coyne, S. M., Manning, J. T., Ringer, L., & Bailey, L. (2007). Directional asymmetry (right-left differences) in digit ratio (2D:4D) predict indirect aggression in women. Personality and Individual Differences, 43(4), 865–872. doi: 10.1016/j.paid.2007.02.010 Coyne, S. M., Nelson, D. A., Graham-Kevan, N., Keister, E., & Grant, D. M. (2010). Mean on the screen: Psychopathy, relationship aggression, and aggression in the media. Personality and Individual Differences, 48(3), 288–293. doi: 10.1016/j.paid.2009.10.018 Coyne, S. M., Nelson, D. A., Graham-Kevan, N., Tew, E., Meng, K. N., & Olsen, J. A. (2011). Media depictions of physical and relational aggression: Connections with aggression in young adults' romantic relationships. Aggressive Behavior, 37, 56–62. doi: 10.1002/ab.20372 Coyne, S. M., Nelson, D. A., Lawton, F., Haslam, S., Rooney, L., Titterington, L.,…Ogunlaja, L. (2008). The effects of viewing physical and relational aggression in the media: Evidence for a cross-over effect. Journal of Experimental Social Psychology, 44(6), 1551–1554. doi: 10.1016/j.jesp.2008.06.006 Coyne, S. M., & Thomas, T. J. (2008). Psychopathy, aggression, and cheating behavior: A test of the Cheater-Hawk hypothesis. Personality and Individual Differences, 44(5), 1105–1115. doi: 10.1016/j.paid.2007.11.002 Crain, M. M., Finch, C. L., & Foster, S. L. (2005). The relevance of the social information processing model for understanding relational aggression in girls. Merrill-Palmer Quarterly, 51(2), 213–249. doi: 10.1353/mpq.2005.0010 Crapanzano, A. M., Frick, P. J., & Terranova, A. M. (2010). Patterns of physical and relational aggression in a school-based sample of boys and girls. Journal of Abnormal Child Psychology, 38(4), 433–445. doi: 10.1007/s10802–009–9376–3 Crick, N. R. (1995). Relational aggression: The role of intent attributions, feelings of distress, and provocation type. Development and Psychopathology, 7(2), 313–322. doi: 10.1017/S0954579400006520

Crick, N. R. (1996). The role of overt aggression, relational aggression, and prosocial behavior in the prediction of children's future social adjustment. Child Development, 67(5), 2317–2327. doi: 10.2307/1131625 Crick, N. R., Bigbee, M. A., & Howes, C. (1996). Gender differences in children's normative beliefs about aggression: How do I hurt thee? Let me count the ways. Child Development, 67(3), 1003–1014. doi: 10.2307/1131876 Crick, N. R., Casas, J. F., & Mosher, M. (1997). Relational and overt aggression in preschool. Developmental Psychology, 33(4), 579–588. doi: 10.1037/0012–1649.33.4.579 Crick, N. R., & Dodge, K. A. (1994). A review and reformulation of social informationprocessing mechanisms in children's social adjustment. Psychological Bulletin, 115, 74–101. doi: 10.1037/0033–2909.115.1.74 Crick, N. R., & Dodge, K. A. (1996). Social information-processing mechanisms in reactive and proactive aggression. Child Development, 67(3), 993–1002. doi: 10.2307/1131875 Crick, N. R., & Dodge, K. A. (1999). “Superiority” is in the eye of the beholder: A comment on Sutton, Smith and Swettenham. Social Development, 8, 128–131. doi: 10.1111/1467– 9507.00084 Crick, N. R., & Grotpeter, J. K. (1995). Relational aggression, gender, and socialpsychological adjustment. Child Development, 66(3), 710–722. doi: 10.2307/1131945 Crick, N. R., & Grotpeter, J. K. (1996). Children's treatment by peers: Victims of relational and overt aggression. Development and Psychopathology, 8(2), 367–380. doi: 10.1017/S0954579400007148 Crick, N. R., Grotpeter, J. K., & Bigbee, M. A. (2002). Relationally and physically aggressive children's intent attributions and feelings of distress for relational and instrumental peer provocations. Child Development, 73(4), 1134–1142. doi: 10.1111/1467–8624.00462 Crick, N. R., Murray-Close, D., Marks, P. E. L., & Mohajeri-Nelson, N. (2009). Aggression and peer relationships in school-age children: Relational and physical aggression in group and dyadic contexts. In K. H. Rubin, W. M. Bukowski, & B. Laursen (Eds.), Handbook of peer interactions, relationships, and groups (pp. 287–302). New York, NY: Guilford Press. Crick, N. R., Murray-Close, D., & Woods, K. (2005). Borderline personality features in childhood: A short-term longitudinal study. Development and Psychopathology, 17(4), 1051– 1070. doi: 10.1017/S0954579405050492 Crick, N. R., & Nelson, D. A. (2002). Relational and physical victimization within friendships: Nobody told me there'd be friends like these. Journal of Abnormal Child Psychology, 30(6), 599–607. doi: 10.1023/A:1020811714064 Crick, N. R., Nelson, D. A., Morales, J. R., Cullerton-Sen, C., Casas, J. F., & Hickman, S. E.

(2001). Relational victimization in childhood and adolescence: I hurt you through the grapevine. In J. Juvonen & S. Graham (Eds.), Peer harassment in school: The plight of the vulnerable and the victimized (pp. 196–214). New York, NY: Guilford Press. Crick, N. R., Ostrov, J. M., Burr, J. E., Cullerton-Sen, C., Jansen-Yeh, E., & Ralston, P. (2006). A longitudinal study of relational and physical aggression in preschool. Journal of Applied Developmental Psychology, 27(3), 254–268. doi: 10.1016/j.appdev.2006.02.006 Crick, N. R., Ostrov, J. M., & Kawabata, Y. (2007). Relational aggression and gender: An overview. In D. J. Flannery, A. T. Vazsonyi, & I. D. Waldman (Eds.), The Cambridge handbook of violent behavior and aggression (pp. 245–259). New York, NY: Cambridge University. Crick, N. R., Ostrov, J. M., & Werner, N. E. (2006). A longitudinal study of relational aggression, physical aggression, and children's social-psychological adjustment. Journal of Abnormal Child Psychology, 34(2), 131–142. doi: 10.1007/s10802–005–9009–4 Crick, N. R., & Werner, N. E. (1998). Response decision processes in relational and overt aggression. Child Development, 69(6), 1630–1639. doi: 10.2307/1132136 Crick, N. R., & Zahn-Waxler, C. (2003). The development of psychopathology in females and males: Current progress and future challenges. Development and Psychopathology, 15(3), 719–742. doi: 10.1017/S095457940300035X Criss, M. M., Lee, T. K., Morris, A. S., Cui, L., Bosler, C. D., Shreffler, K. M., & Silk, J. S. (2015). Link between monitoring behavior and adolescent adjustment: An analysis of direct and indirect effects. Journal of Child and Family Studies, 24, 668–678. doi: 10.1007/s10826–013–9877–0 Cullerton-Sen, C., Cassidy, A. R., Murray-Close, D., Cicchetti, D., Crick, N. R., & Rogosch, F. A. (2008). Childhood maltreatment and the development of relational and physical aggression: The importance of a gender-informed approach. Child Development, 79(6), 1736–1751. doi: 10.1111/j.1467–8624.2008.01222.x Cummings, E. M. (1987). Coping with background anger in early childhood. Child Development, 58(4), 976–984. doi: 10.2307/1130538 Cummings, E. M. (1994). Marital conflict and children's functioning. Social Development, 3, 16–36. Czar, K. A., Dahlen, E. R., Bullock, E. E., & Nicholson, B. C. (2011). Psychopathic personality traits in relational aggression among young adults. Aggressive Behavior, 37(2), 207–214. doi: 10.1002/ab.20381 Davidson, R. J., Putnam, K. M., & Larson, C. L. (2000). Dysfunction in the neural circuitry of emotion regulation—A possible prelude to violence. Science, 289(5479), 591–594. doi: 10.1126/science.289.5479.591

Dawes, M., & Xie, H. (2014). The role of popularity goal in early adolescents' behaviors and popularity status. Developmental Psychology, 50(2), 489–497. doi: 10.1037/a0032999 Dijkstra, J. K., Berger, C., & Lindenberg, S. (2011). Do physical and relational aggression explain adolescents' friendship selection? The competing roles of network characteristics, gender, and social status. Aggressive Behavior, 37, 417–429. doi: 10.1002/ab.20402 Dodge, K. A. (1986). A social information processing model of social competence in children. In M. Perlmutter (Ed.), The Minnesota symposium on child psychology (Vol. 18, pp. 77–125). Hillsdale, NJ: Erlbaum. Dodge, K. A., & Tomlin, A. M. (1987). Utilization of self-schemas as a mechanism of interpretational bias in aggressive children. Social Cognition, 5(3), 280–300. doi: 10.1521/soco.1987.5.3.280 Eder, D. (1985). The cycle of popularity: Interpersonal relations among female adolescents. Sociology of Education, 58(3), 154–165. doi: 10.2307/2112416 Egeland, B., Pianta, R., & Ogawa, J. (1996). Early behavior problems: Pathways to mental disorders in adolescence. Development and Psychopathology, 8(4), 735–749. doi: 10.1017/S0954579400007392 Eisenberger, N. I., Lieberman, M. D., & Williams, K. D. (2003). Does rejection hurt? An fMRI study of social exclusion. Science, 302(5643), 290–292. doi: 10.1126/science.1089134 Ellis, W. E., Crooks, C. V., & Wolfe, D. A. (2009). Relational aggression in peer and dating relationships: Links to psychological and behavioral adjustment. Social Development, 18(2), 253–269. doi: 10.1111/j.1467–9507.2008.00468.x Espelage, D. L., Holt, M. K., & Henkel, R. R. (2003). Examination of peer-group contextual effects on aggression during early adolescence. Child Development, 74, 205–220. doi: 10.1111/1467–8624.00531 Estrem, T. L. (2005). Relational and physical aggression among preschoolers: The effect of language skills and gender. Early Education & Development, 16(2), 207–231. doi: 10.1207/s15566935eed1602_6 Fanger, S. M., Frankel, L. A., & Hazen, N. (2012). Peer exclusion in preschool children's play: Naturalistic observations in a playground setting. Merrill-Palmer Quarterly: Journal of Developmental Psychology, 58(2), 224–254. doi: 10.1353/mpq.2012.0007 Feshbach, N. D. (1969). Sex differences in children's modes of aggressive responses toward outsiders. Merrill-Palmer Quarterly, 15(3), 249–259. Fite, P. J., Stauffacher, K., Ostrov, J. M., & Colder, C. R. (2008). Replication and extension of the Little et al.'s (2003) forms and functions of aggression measure. International Journal of Behavioral Development, 32(3), 238–242. doi: 10.1177/0165025408089273

Fite, P. J., Stoppelbein, L., Greening, L., & Gaertner, A. E. (2009). Further validation of a measure of proactive and reactive aggression within a clinical child population. Child Psychiatry and Human Development, 40(3), 367–382. doi: 10.1007/s10578–009–0132–2 Fite, P. J., Stoppelbein, L., Greening, L., & Preddy, T. M. (2011). Associations between relational aggression, depression, and suicidal ideation in a child psychiatric inpatient sample. Child Psychiatry and Human Development, 42(6), 666–678. doi: 10.1007/s10578–011– 0243–4 Flanagan, T., Iarocci, G., D'Arrisso, A., Mandour, T., Tootoosis, C., Robinson, S., & Burack, J. A. (2011). Reduced ratings of physical and relational aggression for youths with a strong cultural identity: Evidence from the Naskapi people. Journal of Adolescent Health, 49(2), 155–159. doi: 10.1016/j.jadohealth.2010.11.245 Flay, B. R., Biglan, A., Boruch, R. F., Castro, F. G., Gottfredson, D., Kellam, S.,…Ji, P. (2005). Standards of evidence: Criteria for efficacy, effectiveness and dissemination. Prevention Science, 6(3), 151–175. doi: 10.1007/s11121–005–5553-y Forbes, G., Zhang, X., Doroszewicz, K., & Haas, K. (2009). Relationships between individualism-collectivism, gender, and direct or indirect aggression: A study in China, Poland, and the U.S. Aggressive Behavior, 35, 24–30. doi: 10.1002/ab.20292 French, D. C., Jansen, E. A., & Pidada, S. (2002). United States and Indonesian children's and adolescents' reports of relational aggression by disliked peers. Child Development, 73(4), 1143–1150. doi: 10.1111/1467–8624.00463 Frick, P. J., & Morris, A. S. (2004). Temperament and developmental pathways to conduct problems. Journal of Clinical Child and Adolescent Psychology, 33, 54–68. doi: 10.1207/S15374424JCCP3301_6 Gaertner, A. E., Rathert, J. L., Fite, P. J., Vitulano, M., Wynn, P. T., & Harber, J. (2010). Sources of parental knowledge as moderators of the relation between parental psychological control and relational and physical/verbal aggression. Journal of Child and Family Studies, 19(5), 607–616. doi: 10.1007/s10826–009–9345-z Galen, B. R., & Underwood, M. K. (1997). A developmental investigation of social aggression among children. Developmental Psychology, 33(4), 589–600. doi: 10.1037/0012– 1649.33.4.589 Garandeau, C. F., Ahn, H. J., & Rodkin, P. C. (2011). The social status of aggressive students across contexts: The role of classroom status hierarchy, academic achievement, and grade. Developmental Psychology, 47(6), 1699–1710. doi: 10.1037/a0025271 Geiger, T., & Crick, N. R. (2001). A developmental psychopathology perspective on vulnerability to personality disorders. In R. Ingram & J. M. Price (Eds.), Vulnerability to psychopathology: Risk across the life span (pp. 57–102). New York, NY: Guilford Press.

Gentile, D. A., Coyne, S., & Walsh, D. A. (2011). Media violence, physical aggression, and relational aggression in school age children: A short-term longitudinal study. Aggressive Behavior, 37(2), 193–206. doi: 10.1002/ab.20380 Gentile, D. A., Mathieson, L. C., & Crick, N. R. (2011). Media violence associations with the form and function of aggression among elementary school children. Social Development, 20(2), 213–232. doi: 10.1111/j.1467–9507.2010.00577.x Gladden, R. M., Vivolo-Kantor, A. M., Hamburger, M. E., & Lumpkin, C. D. (2014). Bullying Surveillance Among Youths: Uniform Definitions for Public Health and Recommended Data Elements, Version 1.0. Atlanta, GA; National Center for Injury Prevention and Control, Centers for Disease Control and Prevention and U.S. Department of Education. Retrieved from http://www.cdc.gov/violenceprevention/pdf/bullying-definitions-final-a.pdf Godleski, S. A., & Ostrov, J. M. (2010). Relational aggression and hostile attribution biases: Testing multiple statistical methods and models. Journal of Abnormal Child Psychology, 38(4), 447–458. doi: 10.1007/s10802–010–9391–4 Godleski, S. A., Ostrov, J. M., Houston, R. J., & Schlienz, N. J. (2010). Hostile attribution biases for relationally provocative situations and event-related potentials. International Journal of Psychophysiology, 76, 25–33. doi: 10.1016/j.ijpsycho.2010.01.010 Goldstein, S. E. (2011). Relational aggression in young adults' friendships and romantic relationships. Personal Relationships, 18(4), 645–656. doi: 10.1111/j.1475– 6811.2010.01329.x Goldstein, S. E., Chesir-Teran, D., & McFaul, A. (2008). Profiles and correlates of relational aggression in young adults' romantic relationships. Journal of Youth and Adolescence, 37(3), 251–265. doi: 10.1007/s10964–007–9255–6 Goldstein, S. E., & Tisak, M. S. (2004). Adolescents' outcome expectancies about relational aggression within acquaintanceships, friendships, and dating relationships. Journal of Adolescence, 27(3), 283–302. doi: 10.1016/j.adolescence.2003.11.007 Goldstein, S. E., & Tisak, M. S. (2010). Adolescents' social reasoning about relational aggression. Journal of Child and Family Studies, 19(4), 471–482. doi: 10.1007/s10826–009– 9319–1 Goldstein, S. E., Tisak, M. S., & Boxer, P. (2002). Preschoolers' normative and prescriptive judgments about relational and overt aggression. Early Education and Development, 13, 23– 29. doi: 10.1207/s15566935eed1301_2 Golmaryami, F. N., & Barry, C. T. (2010). The associations of self-reported and peer-reported relational aggression with narcissism and self-esteem among adolescents in a residential setting. Journal of Clinical Child and Adolescent Psychology, 39, 128–133. doi: 10.1080/15374410903401203

Gomes, M. M., Davis, B. L., Baker, S. R., & Servonsky, E. J. (2009). Correlation of the experience of peer relational aggression victimization and depression among African American adolescent females. Journal of Child and Adolescent Psychiatric Nursing, 22(4), 175–181. doi: 10.1111/j.1744–6171.2009.00196.x Gouze, K. R. (1987). Attention and social problem solving as correlates of aggression in preschool males. Journal of Abnormal Child Psychology, 15(2), 181–197. doi: 10.1007/BF00916348 Gower, A. L., & Crick, N. R. (2011). Baseline autonomic nervous system arousal and physical and relational aggression in preschool: The moderating role of effortful control. International Journal of Psychophysiology, 81(3), 142–151. doi: 10.1016/j.ijpsycho.2011.06.001 Gower, A. L., Shlafer, R. J., Polan, J., McRee, A.-L., McMorris, B. J., Pettingell, S. L., & Sieving, R. E. (2014). Brief report: Associations between adolescent girls' social–emotional intelligence and violence perpetration. Journal of Adolescence, 37(1), 67–71. doi: 10.1016/j.adolescence.2013.10.012 Gros, D. F., Gros, K. S., & Simms, L. J. (2010). Relations between anxiety symptoms and relational aggression and victimization in emerging adults. Cognitive Therapy and Research, 34(2), 134–143. doi: 10.1007/s10608–009–9236-z Grotpeter, J. K., & Crick, N. R. (1996). Relational aggression, overt aggression, and friendship. Child Development, 67(5), 2328–2338. doi: 10.2307/1131626 Hanish, L. D., & Guerra, N. G. (2002). A longitudinal analysis of patterns of adjustment following peer victimization. Development and Psychopathology, 14, 69–89. doi: 10.1017/S0954579402001049 Hart, C. H., Nelson, D. A., Robinson, C. C., Olsen, S. F., & McNeilly-Choque, M. K. (1998). Overt and relational aggression in Russian nursery-school-age children: Parenting style and marital linkages. Developmental Psychology, 34(4), 687–697. doi: 10.1037/0012– 1649.34.4.687 Hawley, P. H. (2003a). Prosocial and coercive configurations of resource control in early adolescence: A case for the well-adapted Machiavellian. Merrill Palmer Quarterly, 49(3), 279–309. doi: 10.1353/mpq.2003.0013 Hawley, P. H. (2003b). Strategies of control, aggression, and morality in preschoolers: An evolutionary perspective. Journal of Experimental Child Psychology, 85(3), 213–235. doi: 10.1016/S0022–0965(03)00073–0 Hawley, P. H., Little, T. D., & Card, N. A. (2007). The allure of a mean friend: Relationship quality and processes of aggressive adolescents with prosocial skills. International Journal of Behavioral Development, 31(2), 170–180. doi: 10.1177/0165025407074630 Hawley, P. H., Little, T. D., & Card, N. A. (2008). The myth of the alpha male: A new look at

dominance-related beliefs and behaviors among adolescent males and females. International Journal of Behavioral Development, 32, 76–88. doi: 10.1177/0165025407084054 Hemphill, S. A., Kotevski, A., Herrenkohl, T. I., Toumbourou, J. W., Carlin, J. B., Catalano, R. F., & Patton, G. C. (2010). Pubertal stage and the prevalence of violence and social/relational aggression. Pediatrics, 126, 298–305. doi: 10.1542/peds.2009–0574 Hemphill, S. A., Kotevski, A., Tollit, M., Smith, R., Herrenkohl, T. I., Toumbourou, J. W., & Catalano, R. F. (2012). Longitudinal predictors of cyber and traditional bullying perpetration in Australian secondary school students. Journal of Adolescent Health, 51, 59–65. doi: 10.1016/j.jadohealth.2011.11.019 Hoglund, W., Hosan, N. E., & Leadbeater, B. J. (2012). Using your WITS: A 6-year follow-up of a peer victimization prevention program. School Psychology Review, 41(2), 193–214. Hoglund, W. L., & Leadbeater, B. J. (2007). Managing threat: Do social-cognitive processes mediate the link between peer victimization and adjustment problems in early adolescence? Journal of Research on Adolescence, 17(3), 525–540. doi: 10.1111/j.1532– 7795.2007.00533.x Hubbard, J. A., Smithmyer, C. M., Ramsden, S. R., Parker, E. H., Flanagan, K. D., Dearing, K. F.,…Simons, R. F. (2002). Observational, physiological, and self–report measures of children's anger: Relations to reactive versus proactive aggression. Child Development, 73(4), 1101–1118. doi: 10.1111/1467–8624.00460 Huesmann, L. R. (1986). Psychological processes promoting the relation between exposure to media violence and aggressive behavior by the viewer. Journal of Social Issues, 42(3), 125– 139. doi: 10.1111/j.1540–4560.1986.tb00246.x Huesmann, L. R., Moise-Titus, J., Podolski, C.-L., & Eron, L. D. (2003). Longitudinal relations between children's exposure to TV violence and their aggressive and violent behavior in young adulthood: 1977–1992. Developmental Psychology, 39(2), 201–221. doi: 10.1037/0012– 1649.39.2.201 Izard, C. E., Youngstrom, E. A., Fine, S. E., Mostow, A. J., & Trentacosta, C. J. (2006). Emotions and developmental psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Theory and method (Vol. 1, 2nd ed., pp. 244–292). Hoboken, NJ: Wiley. Johnson, D. R., & Foster, S. L. (2005). The relationship between relational aggression in Kindergarten children and friendship stability, mutuality, and peer liking. Early Education and Development, 16(2), 141–160. doi: 10.1207/s15566935eed1602_3 Kamper, K. E., & Ostrov, J. M. (2013). Relational aggression in middle childhood predicting adolescent social-psychological adjustment: The role of friendship quality. Journal of Clinical Child and Adolescent Psychology, 42(6), 855–862. doi: 10.1080/15374416.2013.844595

Kärnä, A., Voeten, M., Little, T. D., Alanen, E., Poskiparta, E., & Salmivalli, C. (2013). Effectiveness of the KiVa antibullying program: Grades 1–3 and 7–9. Journal of Educational Psychology, 105(2), 533–551. doi: 10.1037/a0030417 Kärnä, A., Voeten, M., Little, T., Poskiparta, E., Kaljonen, A., & Salmivalli, C. (2011). A large-scale evaluation of the KiVa antibullying program: Grades 4–6. Child Development, 82, 311–330. doi: 10.1111/j.1467–8624.2010.01557.x Karriker-Jaffe, K. J., Foshee, V. A., Ennett, S. T., & Suchindran, C. (2008). The development of aggression during adolescence: Sex differences in trajectories of physical and social aggression among youth in rural areas. Journal of Abnormal Child Psychology, 36(8), 1227– 1236. doi: 10.1007/s10802–008–9245–5. Karriker-Jaffe, K. J., Foshee, V. A., Ennett, S. T., & Suchindran, C. (2013). Associations of neighborhood and family factors with trajectories of physical and social aggression during adolescence. Journal of Youth and Adolescence, 42(6), 861–877. doi: 10.1007/s10964–012– 9832–1 Kaukiainen, A., Björkqvist, K., Lagerspetz, K., Österman, K., Salmivalli, C., Rothberg, S., & Ahlbom, A. (1999). The relationships between social intelligence, empathy, and three types of aggression. Aggressive Behavior, 25(2), 81–89. doi: 10.1002/(SICI)1098–2337(1999)25:2 Kawabata, Y., Alink, L. R. A., Tseng, W.-L., van IJzendoorn, M. H., & Crick, N. R. (2011). Maternal and paternal parenting styles associated with relational aggression in children and adolescents: A conceptual analysis and meta-analytic review. Developmental Review, 31(4), 240–278. doi: 10.1016/j.dr.2011.08.001 Kawabata, Y., & Crick, N. R. (2013). Relational and physical aggression, peer victimization, and adjustment problems in Asian American and European American children. Asian American Journal of Psychology, 4(3), 211–216. doi: 10.1037/a0031205 Kawabata, Y., Crick, N. R., & Hamaguchi, Y. (2010). The role of culture in relational aggression: Associations with social-psychological adjustment problems in Japanese and US school-aged children. International Journal of Behavioral Development, 34(4), 354–362. doi: 10.1177/0165025409339151 Kawabata, Y., Tseng, W.-L., Murray-Close, D., & Crick, N. R. (2012). Developmental trajectories of Chinese children's relational and physical aggression: Associations with socialpsychological adjustment problems. Journal of Abnormal Child Psychology, 40(7), 1087– 1097. doi: 10.1007/s10802–012–9633–8 Keenan, K., Coyne, C., & Lahey, B. B. (2008). Should relational aggression be included in DSM-V? Journal of the American Academy of Child and Adolescent Psychiatry, 47, 86–93. doi: 10.1097/chi.0b013e31815a56b8 Keenan, K., & Shaw, D. (1997). Developmental and social influences on young girls' early

problem behavior. Psychological Bulletin, 121, 95–113. doi: 10.1037/0033–2909.121.1.95 Kerig, P. K., & Stellwagen, K. K. (2010). Roles of callous-unemotional traits, narcissism, and Machiavellianism in childhood aggression. Journal of Psychopathology and Behavioral Assessment, 32(3), 343–352. doi: 10.1007/s10862–009–9168–7 Kikas, E., Peets, K., Tropp, K., & Hinn, M. (2009). Associations between verbal reasoning, normative beliefs about aggression, and different forms of aggression. Journal of Research on Adolescence, 19, 137–149. doi: 10.1111/j.1532–7795.2009.00586.x Kim-Cohen, J., Caspi, A., Taylor, A., Williams, B., Newcombe, R., Craig, I. W., & Moffitt, T. E. (2006). MAOA, maltreatment, and gene-environment interaction predicting children's mental health: New evidence and a meta-analysis. Molecular Psychiatry, 11(10), 903–913. doi: 10.1038/sj.mp.4001851 Kistner, J., Counts-Allan, C., Dunkel, S., Drew, C. H., David-Ferdon, C., & Lopez, C. (2010). Sex differences in relational and overt aggression in the late elementary school years. Aggressive Behavior, 36, 282–291. doi: 10.1002/ab.20350 Kliewer, W., Dibble, A. E., Goodman, K. L., & Sullivan, T. N. (2012). Physiological correlates of peer victimization and aggression in African American urban adolescents. Development and Psychopathology, 24(2), 637–650. doi: 10.1017/S0954579412000211 Kuppens, S., Grietens, H., Onghena, P., & Michiels, D. (2009a). Associations between parental control and children's overt and relational aggression. British Journal of Developmental Psychology, 27(3), 607–623. doi: 10.1348/026151008X345591 Kuppens, S., Grietens, H., Onghena, P., & Michiels, D. (2009b). A longitudinal study of childhood social behaviour: Inter-informant agreement, inter-context agreement, and social preference linkages. Journal of Social and Personal Relationships, 26(6–7), 769–792. doi: 10.1177/0265407509347929 Kuppens, S., Grietens, H., Onghena, P., & Michiels, D. (2009c). Relations between parental psychological control and childhood relational aggression: Reciprocal in nature? Journal of Clinical Child and Adolescent Psychology, 38, 117–131. doi: 10.1080/15374410802575354 Kuppens, S., Grietens, H., Onghena, P., Michiels, D., & Subramanian, S. V. (2008). Individual and classroom variables associated with relational aggression in elementary-school aged children: A multilevel analysis. Journal of School Psychology, 46(6), 639–660. doi: 10.1016/j.jsp.2008.06.005 Kuppens, S., Laurent, L., Heyvaert, M., & Onghena, P. (2013). Associations between parental psychological control and relational aggression in children and adolescents: A multilevel and sequential meta-analysis. Developmental Psychology, 49(9), 1697–1712. doi: 10.1037/a0030740 LaFontana, K. M., & Cillessen, A. H. N. (2010). Developmental changes in the priority of

perceived status in childhood and adolescence. Social Development, 19, 130–147. doi: 10.1111/j.1467–9507.2008.00522.x Lagerspetz, K. M. J., & Björkqvist, K. (1994). Indirect aggression in boys and girls. In L. R. Huesmann (Ed.), Aggressive behavior: Current perspectives (pp. 131–150). New York, NY: Plenum Press. Lagerspetz, K. M. J., Björkqvist, K., & Peltonen, T. (1988). Is indirect aggression typical of females? Gender differences in aggressiveness in 11- to 12-year-old children. Aggressive Behavior, 14(6), 403–414. doi: 10.1002/1098–2337(1988)14:6 Landau, S. F., Björkqvist, K., Lagerspetz, K. M. J., Österman, K., & Gideon, L. (2002). The effect of religiosity and ethnic origin on direct and indirect aggression among males and females: Some Israeli findings. Aggressive Behavior, 28(4), 281–298. doi: 10.1002/ab.80006 Lansford, J. E., Miller, S., Malone, P. S., Costanzo, P. R., Grimes, C., & Putallaz, M. (2009). Social network centrality and leadership status: Links with problem behaviors and tests of gender differences. Merrill-Palmer Quarterly, 55(1), 1–25. doi: 10.1353/mpq.0.0014 Lansford, J. E., Skinner, A. T., Sorbring, E., Di Giunta, L., Deater-Deckard, K., Dodge, K. A., …Chang, L. (2012). Boys' and girls' relational and physical aggression in nine countries. Aggressive Behavior, 38(4), 298–308. doi: 10.1002/ab.21433 Lau, K. S. L., & Marsee, M. A. (2013). Exploring narcissism, psychopathy, and Machiavellianism in youth: Examination of associations with antisocial behavior and aggression. Journal of Child and Family Studies, 22(3), 355–367. doi: 10.1007/s10826–012– 9586–0 Leadbeater, B. J., Banister, E. M., Ellis, W. E., & Yeung, R. (2008). Victimization and relational aggression in adolescent romantic relationships: The influence of parental and peer behaviors, and individual adjustment. Journal of Youth and Adolescence, 37(3), 359–372. doi: 10.1007/s10964–007–9269–0 Leadbeater, B. J., Boone, E. M., Sangster, N. A., & Mathieson, L. C. (2006). Sex differences in the personal costs and benefits of relational and physical aggression in high school. Aggressive Behavior, 32(4), 409–419. doi: 10.1002/ab.20139 Leadbeater, B., Hoglund, W., & Woods, T. (2003). Changing contexts? The effects of a primary prevention program on classroom levels of peer relational and physical victimization. Journal of Community Psychology, 31, 397–418. doi: 10.1002/jcop.10057 Lease, A. M., Kennedy, C. A., & Axelrod, J. L. (2002). Children's social constructions of popularity. Social Development, 11, 87–109. doi: 10.1111/1467–9507.00188 Leavitt, C. E., Nelson, D. A., Coyne, S. M., & Hart, C. H. (2013). Adolescent disclosure and concealment: Longitudinal and concurrent associations with aggression. Aggressive Behavior, 39, 335–345. doi: 10.1002/ab.21488

Leff, S. S., Baker, C. N., Waasdorp, T. E., Vaughn, N. A., Bevans, K., Thomas, N. A.,… Monopoli, W. J. (2014). Social cognitions, distress, and leadership self-efficacy: Associations with aggression for high-risk minority youth. Development and Psychopathology, 26, 759– 772. doi: 10.1017/S0954579414000376 Leff, S. S., Gullan, R. L., Paskewich, B. S., Abdul-Kabir, S., Jawad, A. F., Grossman, M.,… Power, T. J. (2009). An initial evaluation of a culturally-adapted social problem solving and relational aggression prevention program for urban African American relationally aggressive girls. Journal of Prevention and Intervention in the Community, 37(4), 260–274. doi: 10.1080/10852350903196274 Leff, S. S., Waasdorp, T. E. & Crick, N. R. (2010). A review of existing relational aggression programs: Strengths, limitations, and future directions. School Psychology Review, 39(4), 508–535. Li, Y., Putallaz, M., & Su, Y. (2011). Interparental conflict styles and parenting behaviors: Associations with overt and relational aggression among Chinese children. Merrill-Palmer Quarterly, 57(4), 402–428. doi: 10.1353/mpq.2011.0017 Li, Y., Wang, M., Wang, C., & Shi, J. (2010). Individualism, collectivism, and Chinese adolescents' aggression: Intracultural variations. Aggressive Behavior, 36(3), 187–194. doi: 10.1002/ab.20341 Li, Y., & Wright, M. F. (2014). Adolescents' social status goals: Relationships to social status insecurity, aggression, and prosocial behavior. Journal of Youth and Adolescence, 43(1), 146– 160. doi: 10.1007/s10964–013–9939-z Linder, J. R., Crick, N. R., & Collins, W. A. (2002). Relational aggression and victimization in young adults' romantic relationships: Associations with perceptions of parent, peer, and romantic relationship quality. Social Development, 11, 69–86. doi: 10.1111/1467–9507.00187 Linder, J. R., & Werner, N. E. (2012). Relationally aggressive media exposure and children's normative beliefs: Does parental mediation matter? Family Relations: An Interdisciplinary Journal of Applied Family Studies, 61(3), 488–500. doi: 10.111/j.1741–3729.2012.00707.x Little, T. D., Jones, S. M., Henrich, C. C., & Hawley, P. H. (2003). Disentangling the “whys” from the “whats” of aggressive behaviour. International Journal of Behavioral Development, 27(2), 122–133. doi: 10.1080/01650250244000128 Loeber, R., Pardini, D. A., Hipwell, A., Stouthamer-Loeber, M., Keenan, K., & Sembower, M. A. (2009). Are there stable factors in preadolescent girls' externalizing behaviors? Journal of Abnormal Child Psychology, 37(6), 777–791. doi: 10.1007/s10802–009–9320–6 Loudin, J. L., Loukas, A., & Robinson, S. (2003). Relational aggression in college students: Examining the roles of social anxiety and empathy. Aggressive Behavior, 29(5), 430–439. doi: 10.1002/ab.10039

Low, S., Polanin, J. R., & Espelage, D. L. (2013). The role of social networks in physical and relational aggression among young adolescents. Journal of Youth and Adolescence, 42(7), 1078–1089. doi: 10.1007/s10964–013–9933–5 MacBrayer, E. K., Milich, R., & Hundley, M. (2003). Attributional biases in aggressive children and their mothers. Journal of Abnormal Psychology, 112(4), 698–708. doi: 10.1037/0021–843X.112.4.598 Maccoby, E. E., & Jacklin, C. N. (1974). The psychology of sex differences. Stanford, CA: Stanford University Press. MacEvoy, J. P., & Leff, S. S. (2012). Children's sympathy for peers who are the targets of peer aggression. Journal of Abnormal Child Psychology, 40(4), 1137–1148. doi: 10.1007/s10802– 012–9636–5 Marsee, M. A., & Frick, P. J. (2007). Exploring the cognitive and emotional correlates to proactive and reactive aggression in a sample of detained girls. Journal of Abnormal Child Psychology, 35(6), 969–981. doi: 10.1007/s10802–007–9147-y Marsee, M. A., Frick, P. J., Barry, C. T., Kimonis, E. R., Centifanti, L. C. M., & Aucoin, K. J. (2014). Profiles of the forms and functions of self-reported aggression in three adolescent samples. Development and Psychopathology, 26, 705–720. doi: 10.1017/S0954579414000339 Marsee, M. A., Silverthorn, P., & Frick, P. J. (2005). The association of psychopathic traits with aggression and delinquency in non-referred boys and girls. Behavioral Sciences & the Law, 23(6), 803–817. doi: 10.1002/bsl.662 Marsee, M. A., Weems, C. F., & Taylor, L. K. (2008). Exploring the association between aggression and anxiety in youth: A look at aggressive subtypes, gender, and social cognition. Journal of Child and Family Studies, 17, 154–168. doi: 10.1007/s10826–007–9154–1 Martin, C. L., & Halverson, C. F. (1981). A schematic processing model of sex typing and stereotyping in children. Child Development, 52(4), 1119–1134. doi: 10.2307/1129498 Martins, N. (2013). Televised relational and physical aggression and children's hostile intent attributions. Journal of Experimental Child Psychology, 116(4), 945–952. doi: 10.1016/j.jecp.2013.05.006 Masten, A. S., & Coatsworth, J. D. (1998). The development of competence in favorable and unfavorable environments: Lessons from research on successful children. American Psychologist, 53(2), 205–220. doi: 10.1037/0003–066X.53.2.205 Masten, A. S., Roisman, G. I., Long, J. D., Burt, K. B., Obradović, J., Riley, J. R.,…Tellegen, A. (2005). Developmental cascades: Linking academic achievement and externalizing and internalizing symptoms over 20 years. Developmental Psychology, 41(5), 733–746. doi: 10.1037/0012–1649.41.5.733

Mathieson, L. C., & Crick, N. R. (2010). Reactive and proactive subtypes of relational and physical aggression in middle childhood: Links to concurrent and longitudinal adjustment. School Psychology Review, 39(4), 601–611. Mathieson, L. C., Klimes-Dougan, B., & Crick, N. R. (2014). Dwelling on it may make it worse: The links between relational victimization, relational aggression, rumination, and depressive symptoms in adolescents. Development and Psychopathology, 26, 735–747. doi: 10.1017/S0954579414000352 Mathieson, L. C., Murray-Close, D., Crick, N. R., Woods, K. E., Zimmer-Gembeck, M., Geiger, T. C., & Morales, J. R. (2011). Hostile intent attributions and relational aggression: The moderating roles of emotional sensitivity, gender, and victimization. Journal of Abnormal Child Psychology, 39(7), 977–987. doi: 10.1007/s10802–011–9515–5 Mayeux, L. M., & Cillessen, A. H. N. (2008). It's not just being popular, it's knowing it, too: The role of self-perceptions of status in the associations between peer status and aggression. Social Development, 17(4), 871–888. doi: 10.1111/j.1467–9507.2008.00474.x McHale, S. M., & Crouter, A. C. (1996). The family contexts of children's sibling relationships. In G. H. Brody (Ed.), Sibling relationships: Their causes and consequences (Vol. 1, pp. 173–195). Norwood, NJ: Ablex. McNamara, K. A., Selig, J. P., & Hawley, P. H. (2010). A typological approach to the study of parenting: Associations between maternal parenting patterns and child behaviour and social reception. Early Child Development and Care, 180(9), 1185–1202. doi: 10.1080/03004430902907574 McNeilly-Choque, M. K., Hart, C. H., Robinson, C. C., Nelson, L. J., & Olsen, S. F. (1996). Overt and relational aggression on the playground: Correspondence among different informants. Journal of Research in Childhood Education, 11(1), 47–67. Merten, D. E. (1997). The meaning of meanness: Popularity, competition and conflict among junior high school girls. Sociology of Education, 70, 175–191. Michiels, D., Grietens, H., Onghena, P., & Kuppens, S. (2008). Parent-child interactions and relational aggression in peer relationships. Developmental Review, 28(4), 522–540. doi: 10.1016/j.dr.2008.08.002 Mikami, A. Y., Lee, S. S., Hinshaw, S. P., & Mullin, B. C. (2008). Relationships between social information processing and aggression among adolescent girls with and without ADHD. Journal of Youth and Adolescence, 37(7), 761–771. doi: 10.1007/s10964–007–9237–8 Miller, J. D., & Lynam, D. R. (2003). Psychopathy and the five-factor model of personality: A replication and extension. Journal of Personality Assessment, 81(2), 168–178. doi: 10.1207/S15327752JPA8102_08 Miller, J. L., Vaillancourt, T., & Boyle, M. H. (2009). Examining the heterotypic continuity of

aggression using teacher reports: Results from a national Canadian study. Social Development, 18, 164–180. doi: 10.1111/j.1467–9507.2008.00480.x Moffitt, T. E., & Caspi, A. (2001). Childhood predictors differentiate life-course persistent and adolescence-limited antisocial pathways among males and females. Development and Psychopathology, 13(2), 355–375. doi: 10.1017/S0954579401002097 Möller, I., & Krahé, B. (2009). Exposure to violent video games and aggression in German adolescents: A longitudinal analysis. Aggressive Behavior, 35, 75–89. doi: 10.1002/ab.20290 Moore, C. C., Shoulberg, E. K., & Murray-Close, D. (2012). The protective role of teacher preference for at-risk children's social status. Aggressive Behavior, 38(6), 481–493. doi: 10.1002/ab.21446 Murray, K. W., Haynie, D. L., Howard, D. E., Cheng, T. L., & Simons-Morton, B. (2010). Perceptions of parenting practices as predictors of aggression in a low-income, urban, predominately African American middle school sample. Journal of School Violence, 9(2), 174–193. doi: 10.1080/15388220903585853 Murray-Close, D. (2011). Autonomic reactivity and romantic relational aggression among female emerging adults: Moderating roles of social and cognitive risk. International Journal of Psychophysiology, 80, 28–35. doi: 10.1016/j.ijpsycho.2011.01.007 Murray-Close, D. (2013a). Psychophysiology of adolescent peer relations I: Theory and research findings. Journal of Research on Adolescence, 23(2), 236–259. doi: 0.1111/j.1532– 7795.2012.00828.x Murray-Close, D. (2013b). Psychophysiology of adolescent peer relations II: Recent advances and future directions. Journal of Research on Adolescence, 23(2), 260–273. doi: 10.1111/j.1532–7795.2012.00831.x Murray-Close, D., & Crick, N. R. (2006). Mutual antipathy involvement: Gender and associations with aggression and victimization. School Psychology Review, 35(3), 472–492. Murray-Close, D., & Crick, N. R. (2007). Gender differences in the association between cardiovascular reactivity and aggressive conduct. International Journal of Psychophysiology, 65(2), 103–113. doi: 10.1016/j.ijpsycho.2007.03.011 Murray-Close, D., Crick, N. R., Tseng, W.-L., Lafko, N., Burrows, C., Pitula, C., & Ralston, P. (2014). Physiological stress reactivity and physical and relational aggression: The moderating roles of victimization, type of stressor, and child gender. Development and Psychopathology, 26, 589–603. doi: 10.1017/S095457941400025X Murray-Close, D., Han, G., Cicchetti, D., Crick, N. R., & Rogosch, F. A. (2008). Neuroendocrine regulation and physical and relational aggression: The moderating roles of child maltreatment and gender. Developmental Psychology, 44(4), 1160–1176. doi: 10.1037/a0012564

Murray-Close, D., Holland, A. S., & Roisman, G. I. (2012). Autonomic arousal and relational aggression in heterosexual dating couples. Personal Relationships, 19(2), 203–218. doi: 10.1111/j.1475–6811.2011.01348.x Murray-Close, D. & Ostrov, J. M. (2009). A longitudinal study of forms and functions of aggressive behavior in early childhood. Child Development, 80(3), 828–842. doi: 10.1111/j.1467–8624.2009.01300.x Murray-Close, D., Ostrov, J. M., & Crick, N. R. (2007). A short-term longitudinal study of growth of relational aggression during middle childhood: Associations with gender, friendship intimacy, and internalizing problems. Development and Psychopathology, 19, 187–203. doi: 10.1017/S0954579407070101 Murray-Close, D., Ostrov, J. M., Nelson, D. A., Crick, N. R., & Coccaro, E. F. (2010). Proactive, reactive, and romantic relational aggression in adulthood: Measurement, predictive validity, gender differences, and association with intermittent explosive disorder. Journal of Psychiatric Research, 44(6), 393–404. doi: 10.1016/j.jpsychires.2009.09.005 Murray-Close, D., & Rellini, A. H. (2012). Cardiovascular reactivity and proactive and reactive relational aggression among women with and without a history of sexual abuse. Biological Psychology, 89, 54–62. doi: 10.1016/j.biopsycho.2011.09.008 Musher-Eizenman, D. R., Boxer, P., Danner, S., Dubow, E. F., Goldstein, S. E., & Heretick, D. M. L. (2004). Social-cognitive mediators of the relation of environmental and emotion regulation factors to children's aggression. Aggressive Behavior, 30(4), 389–409. doi: 10.1002/ab.20078 Neal, J. W., & Cappella, E. (2012). An examination of network position and childhood relational aggression: Integrating resource control and social exchange theories. Aggressive Behavior, 38(2), 126–140. doi: 10.1002/ab.21414 Neiderhiser, J. M., Reiss, D., & Hetherington, E. M. (1996). Genetically informative designs for distinguishing developmental pathways during adolescence: Responsible and antisocial behavior. Development and Psychopathology, 8(4), 779–791. doi: 10.1017/S0954579400007422 Nelson, D. A., & Coyne, S. M. (2009). Children's intent attributions and feelings of distress: Associations with maternal and paternal parenting practices. Journal of Abnormal Child Psychology, 37(2), 223–237. doi: 10.1007/s10802–008–9271–3 Nelson, D. A., Coyne, S. M., Swanson, S. M., Hart, C. H., & Olsen, J. A. (2014). Parenting, relational aggression, and borderline personality features: Associations over time in a Russian longitudinal sample. Development and Psychopathology, 26, 773–87. doi: 10.1017/S0954579414000388 Nelson, D. A., & Crick, N. R. (2002). Parental psychological control: Implications for

childhood physical and relational aggression. In B. K. Barber (Ed.), Intrusive parenting: How psychological control affects children and adolescents (pp. 161–189). Washington, DC: American Psychological Association. Nelson, D. A., Hart, C. H., Yang, C., Olsen, J. A., & Jin, S. (2006). Aversive parenting in China: Associations with child physical and relational aggression. Child Development, 77(3), 554–572. doi: 10.1111/j.1467–8624.2006.00890.x Nelson, D. A., Mitchell, C., & Yang, C. (2008). Intent attributions and aggression: A study of children and their parents. Journal of Abnormal Child Psychology, 36(6), 793–806. doi: 10.1007/s10802–007–9211–7 Nelson, D. A., Robinson, C. C., & Hart, C. H. (2005). Relational and physical aggression of preschool-age children: Peer status linkages across informants. Early Education and Development: Special Issue on Relational Aggression in Early Childhood, 16(2), 115–139. doi: 10.1207/s15566935eed1602_2 Nelson, D. A., Robinson, C. C., Hart, C. H., Albano, A. D., & Marshall, S. J. (2010). Italian preschoolers' peer-status linkages with sociability and subtypes of aggression and victimization. Social Development, 19(4), 698–720. doi: 10.1111/j.1467–9507.2009.00551.x Nelson, D. A., Springer, M. M., Nelson, L. J., & Bean, N. H. (2008). Normative beliefs regarding aggression in emerging adulthood. Social Development, 17(3), 638–660. doi: 10.1111/j.1467–9507.2007.00442.x Nelson, D. A., Yang, C., Coyne, S. M., Olsen, J. A., & Hart, C. H. (2013). Parental psychological control dimensions: Connections with Russian preschoolers' physical and relational aggression. Journal of Applied Developmental Psychology, 34, 1–8. doi: 10.1016/j.appdev.2012.07.003 Ojanen, T., Findley, D., & Fuller, S. (2012). Physical and relational aggression in early adolescence: Associations with narcissism, temperament, and social goals. Aggressive Behavior, 38(2), 99–107. doi: 10.1002/ab.21413 Onishi, A., Kawabata, Y., Kurokawa, M., & Yoshida, T. (2012). A mediating model of relational aggression, narcissistic orientations, guilt feelings, and perceived classroom norms. School Psychology International, 33, 367–390. doi: 10.1177/0143034311421433 Ortiz, J., & Raine, A. (2004). Heart rate level and antisocial behavior in children and adolescents: A meta-analysis. Journal of the American Academy of Child & Adolescent Psychiatry, 43(2), 154–162. doi: 10.1097/00004583–200402000–00010 Orobio de Castro, B., Thomaes, S., & Reijntjes, A. (2015). Using experimental designs to understand the development of peer relations. Journal of Research on Adolescence, 25, 1–13. doi: 10.1111/jora.12103 Orobio de Castro, B., Veerman, J. W., Koops, W., Bosch, J. D., & Monshouwer, H. J. (2002).

Hostile attribution of intent and aggressive behavior: A meta-analysis. Child Development, 73(3), 916–934. doi: 10.1111/1467–8624.00447 Orue, I., & Calvete, E. (2011). Reciprocal relationships between sociometric indices of social status and aggressive behavior in children: Gender differences. Journal of Social and Personal Relationships, 28(7), 963–982. doi: 10.1177/0265407510397982 Österman, K., Björkqvist, K., Lagerspetz, K. M. J., Kaukiainen, A., Huesmann, L. R., & Frączek, A. (1994). Peer and self-estimated aggression and victimization in 8-year-old children from five ethnic groups. Aggressive Behavior, 20(6), 411–428. doi: 10.1002/1098– 2337(1994)20:6 Österman, K., Björkqvist, K., Lagerspetz, K. M. J., Kaukiainen, A., Landau, S. F., Frączek, A., & Caprara, G. V. (1998). Cross-cultural evidence of female indirect aggression. Aggressive Behavior, 24, 1–8. doi: 10.1002/(SICI)1098–2337(1998)24:1 Ostrov, J. M. (2006). Deception and subtypes of aggression during early childhood. Journal of Experimental Child Psychology, 93(4), 322–336. doi: 10.1016/j.jecp.2005.10.004 Ostrov, J. M. (2008). Forms of aggression and peer victimization during early childhood: A short-term longitudinal study. Journal of Abnormal Child Psychology, 36(3), 311–322. doi: 10.1007/s10802–007–9179–3 Ostrov, J. M. (2010). Prospective associations between peer victimization and aggression. Child Development, 81(6), 1670–1677. doi: 10.1111/j.1467–8624.2010.01501.x Ostrov, J. M., & Bishop, C. M. (2008). Preschoolers' aggression and parent-child conflict: A multiinformant and multimethod study. Journal of Experimental Child Psychology, 99(4), 309–322. doi: 10.1016/j.jecp.2008.01.001 Ostrov, J. M., & Crick, N. R. (2007). Forms and functions of aggression in early childhood: A short-term longitudinal study. School Psychology Review, 36, 22–43. Ostrov, J. M., Crick, N. R., & Stauffacher, K. (2006). Relational aggression in sibling and peer relationships during early childhood. Journal of Applied Developmental Psychology, 27(3), 241–253. doi: 10.1016/j.appdev.2006.02.005 Ostrov, J. M., Gentile, D. A., & Crick, N. R. (2006). Media exposure, aggression, and prosocial behavior during early childhood: A longitudinal study. Social Development, 15(4), 612–627. doi: 10.1111/j.1467–9507.2006.00360.x Ostrov, J. M., Gentile, D. A., & Mullins, A. D. (2013). Evaluating the effect of educational media exposure on aggression in early childhood. Journal of Applied Developmental Psychology, 34, 38–44. doi: 10.1016/j.appdev.2012.09.005 Ostrov, J. M., & Godleski, S. A. (2009). Impulsivity-hyperactivity and subtypes of aggression in early childhood: An observational and short-term longitudinal study. European Child and

Adolescent Psychiatry, 18(8), 477–483. doi: 10.1007/s00787–009–0002–2 Ostrov, J. M., & Godleski, S. A. (2010). Toward an integrated gender-linked model of aggression subtypes in early and middle childhood. Psychological Review, 117, 233–242. doi: 10.1037/a0018070 Ostrov, J. M., & Godleski, S. A. (2013). Relational aggression, victimization, and adjustment during middle childhood. Development and Psychopathology, 25(3), 801–815. doi: 10.1017/30954579413000187 Ostrov, J. M., Hart, E. J., Kamper, K. E., & Godleski, S. A. (2011). Relational aggression in women during emerging adulthood: A social process model. Behavioral Sciences & the Law, 29(5), 695–710. doi: 10.1002/bsl.1002 Ostrov, J. M., & Houston, R. J. (2008). The utility of forms and functions of aggression in emerging adulthood: Associations with personality disorder symptomatology. Journal of Youth and Adolescence, 37(9), 1147–1158. doi: 10.1007/s10964–008–9289–4 Ostrov, J. M., Kamper, K. E., Hart, E. J., Godleski, S. A., & Blakely-McClure, S. J. (2014). A gender-balanced approach to the study of peer victimization and aggression subtypes in early childhood. Development and Psychopathology, 26, 575–587. doi: 10.1017/S0954579414000248 Ostrov, J. M., & Keating, C. F. (2004). Gender differences in preschool aggression during free play and structured interactions: An observational study. Social Development, 13(2), 255–277. doi: 10.1111/j.1467–9507.2004.000266.x Ostrov, J. M., Massetti, G. M., Stauffacher, K., Godleski, S. A., Hart, K. C., Karch, K. M.,… Ries, E. E. (2009). An intervention for relational and physical aggression in early childhood: A preliminary study. Early Childhood Research Quarterly, 24, 15–28. doi: 10.1016/j.ecresq.2008.08.002 Ostrov, J. M., Murray-Close, D., Godleski, S. A., & Hart, E. J. (2013). Prospective associations between forms and functions of aggression and social and affective processes during early childhood. Journal of Experimental Child Psychology, 116, 19–36. doi: 10.1016/j.jecp.2012.12.009 Ostrov, J. M., Ries, E. E., Stauffacher, K., Godleski, S. A., & Mullins, A. D. (2008). Relational aggression, physical aggression and deception during early childhood: A multimethod, multi-informant short-term longitudinal study. Journal of Clinical Child and Adolescent Psychology, 37(3), 664–675. doi: 10.1080/15374410802148137 Ostrov, J. M., Woods, K. E., Jansen, E. A., Casas, J. F., & Crick, N. R. (2004). An observational study of delivered and received aggression, gender, and social-psychological adjustment in preschool: “This white crayon doesn't work. . . .” Early Childhood Research Quarterly, 19(2), 355–371. doi: 10.1016/j.ecresq.2004.04.009

Owens, L., Shute, R., & Slee, P. (2000). “Guess what I just heard!”: Indirect aggression among teenage girls in Australia. Aggressive Behavior, 26, 67–83. doi: 10.1002/(SICI)1098– 2337(2000)26:1 Parker, J. G., Rubin, K. H., Price, J. M., & DeRosier, M. E. (1995). Peer relationships, child development, and adjustment: A developmental psychopathology perspective. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Risk, disorder, and adaptation (Vol. 2, 2nd ed., pp. 96–161). Oxford, UK: Wiley. Parkhurst, J. T, & Hopmeyer, A. (1998). Sociometric popularity and peer-perceived popularity: Two distinct dimensions of peer status. Journal of Early Adolescence, 18(2), 125– 144. doi: 10.1177/0272431698018002001 Pashler, H., & Harris, C. R. (2012). Is the replicability crisis overblown? Three arguments examined. Perspectives on Psychological Science, 7(6), 531–536. doi: 10.1177/1745691612463401 Pellegrini, A. D., & Long, J. D. (2003). A sexual selection theory longitudinal analysis of sexual segregation and integration in early adolescence. Journal of Experimental Child Psychology, 85(3), 257–278. doi: 10.1016/S0022–0965(03)00060–2 Peters, E., Cillessen, A. H. N., Riksen-Walraven, J. M., & Haselager, G. J. T. (2010). Best friends' preference and popularity: Associations with aggression and prosocial behavior. International Journal of Behavioral Development, 34(5), 398–405. doi: 10.1177/0165025409343709 Peterson, C. K., Shackman, A. J., & Harmon-Jones, E. (2008). The role of asymmetrical frontal cortical activity in aggression. Psychophysiology, 45, 86–92. doi: 10.1111/j.1469– 8986.2007.00597.x Poteat, V. P., Kimmel, M. S., & Wilchins, R. (2011). The moderating effects of support for violence beliefs on masculine norms, aggression, and homophobic behavior during adolescence. Journal of Research on Adolescence, 21, 434–447. doi: 10.1111/j.1532– 7795.2010.00682.x Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–143. doi: 10.1016/j.biopsycho.2006.06.009 Poulin, F., & Boivin, M. (2000). Reactive and proactive aggression: Evidence of a two-factor model. Psychological Assessment, 12(2), 115–122. doi: 10.1037/1040–3590.12.2.115 Preddy, T. M., & Fite, P. J. (2012). Differential associations between relational and overt aggression and children's psychosocial adjustment. Journal of Psychopathology and Behavioral Assessment, 34(2), 182–190. doi: 10.1007/s10862–011–9274–1 Prinstein, M. J., Boergers, J., & Vernberg, E. M. (2001). Overt and relational aggression in adolescents: Social-psychological adjustment of aggressors and victims. Journal of Clinical

Child Psychology, 30(4), 479–491. doi: 10.1207/S15374424JCCP3004_05 Prinstein, M. J., & Cillessen, A. H. N. (2003). Forms and functions of adolescent peer aggression associated with high levels of peer status. Merrill-Palmer Quarterly, 49(3), 310– 342. doi: 10.1353/mpq.2003.0015 Pronk, R. E., & Zimmer-Gembeck, M. J. (2010). It's “mean,” but what does it mean to adolescents? Relational aggression described by victims, aggressors, and their peers. Journal of Adolescent Research, 25(2), 175–204. doi: 10.1177/0743558409350504 Puckett, M. B., Aikins, J. W., & Cillessen, A. H. N. (2008). Moderators of the association between relational aggression and perceived popularity. Aggressive Behavior, 34(6), 563– 576. doi: 10.1002/ab.20280 Putallaz, M., Grimes, C. L., Foster, K. J., Kupersmidt, J. B., Coie, J. D., & Dearing, K. (2007). Overt and relational aggression and victimization: Multiple perspectives within the school setting. Journal of School Psychology, 45, 523–547. doi: 10.1016/j.jsp.2007.05.003 Quiggle, N. L., Garber, J., Panak, W., & Dodge, K. A. (1992). Social-information processing in aggressive and depressed children. Child Development, 63(6), 1305–1320. doi: 10.2307/1131557 Raine, A. (2002). Biosocial studies of antisocial and violent behavior in children and adults: A review. Journal of Abnormal Child Psychology, 30(4), 311–326. doi: 10.1023/A:1015754122318 Renouf, A., Brendgen, M., Parent, S., Vitaro, F., Zelazo, P. D., Boivin, M.,…Séguin, J. R. (2010). Relations between theory of mind and indirect and physical aggression in kindergarten: Evidence of the moderating role of prosocial behaviors. Social Development, 19(3), 535–555. doi: 10.1111/j.1467–9507.2009.00552.x Rose, A. J., & Swenson, L. P. (2009). Do perceived popular adolescents who aggress against others experience emotional adjustment problems themselves? Developmental Psychology, 45, 868–872. doi: 10.1037/a0015408 Rose, A. J., Swenson, L. P., & Carlson, W. (2004). Friendships of aggressive youth: Considering the influences of being disliked and of being perceived as popular. Journal of Experimental Child Psychology, 88, 25–45. doi: 10.1016/j.jecp.2004.02.005 Rose, A. J., Swenson, L. P., & Waller, E. M. (2004). Overt and relational aggression and perceived popularity: Developmental differences in concurrent and prospective relations. Developmental Psychology, 40(3), 378–387. doi: 10.1037/0012–1649.40.3.378 Rosen, L. H., & Underwood, M. K. (2010). Facial attractiveness as a moderator of the association between social and physical aggression and popularity in adolescents. Journal of School Psychology, 48(4), 313–333. doi: 10.1016/j.jsp.2010.03.001

Rubin, K., Bukowski, W., & Parker, J. G. (2006). Peer interactions, relationships, and groups. In N. Eisenberg, W. Damon, & R. Lerner (Eds.), Handbook of child psychology: Social, emotional, and personality development (Vol. 3, 6th ed., pp. 571–645). Hoboken, NJ: Wiley. Rudolph, K. D., Hammen, C., Burge, D., Lindberg, N., Herzberg, L., & Daley, S. E. (2000). Toward an interpersonal life-stress model of depression: The developmental context of stress generation. Development and Psychopathology, 12(2), 215–234. doi: 10.1017/S0954579400002066 Rudolph, K. D., Troop-Gordon, W., Hessel, E. T., & Schmidt, J. D. (2011). A latent growth curve analysis of early and increasing peer victimization as predictors of mental health across elementary school. Journal of Clinical Child and Adolescent Psychology, 40, 111–122. doi: 10.1080/15374416.2011.533413 Russell, A., Hart, C. H., Robinson, C. C., & Olsen, S. F. (2003). Children's sociable and aggressive behavior with peers: A comparison of the U.S. and Australia, and contributions of temperament and parenting styles. International Journal of Behavioral Development, 27, 74– 86. doi: 10.1080/01650250244000038 Rys, G. S., & Bear, G. G. (1997). Relational aggression and peer relations: Gender and developmental issues. Merrill-Palmer Quarterly, 43, 87–106. Salmivalli, C., Kärnä, A., & Poskiparta, E. (2011). Counteracting bullying in Finland: The KiVa program and its effects on different forms of being bullied. International Journal of Behavioral Development, 35(4), 405–411. doi: 10.1177/0165025411407457 Sameroff, A. J., & MacKenzie, M. J. (2003). Research strategies for capturing transactional models of development: The limits of the possible. Development and Psychopathology, 15(3), 613–640. doi: 10.1017/S0954579403000312 Sánchez-Martín, J. R., Azurmendi, A., Pascual-Sagastizabal, E., Cardas, J., Braza, F., Braza, P.,…Muñoz, J. M. (2011). Androgen levels and anger and impulsivity measures as predictors of physical, verbal and indirect aggression in boys and girls. Psychoneuroendocrinology, 36(5), 750–760. doi: 10.1016/j.psyneuen.2010.10.011 Sandstrom, M. J. (2007). A link between mothers' disciplinary strategies and children's relational aggression. British Journal of Developmental Psychology, 25(3), 399–407. doi: 10.1348/026151006X158753 Sandstrom, M. J., & Cillessen, A. H. N. (2006). Likable versus popular: Distinct implications for adolescent adjustment. International Journal of Behavioral Development, 30(4), 305– 314. doi: 10.1177/0165025406072789 Sandstrom, M. J., & Cillessen, A. H. N. (2010). Life after high school: Adjustment of popular teens in emerging adulthood. Merrill-Palmer Quarterly, 56(4), 474–499. doi: 10.1353/mpq.2010.0000

Schmeelk, K. M., Sylvers, P., & Lilienfeld, S. O. (2008). Trait correlates of relational aggression in a nonclinical sample: DSM-IV personality disorders. Journal of Personality Disorders, 22(3), 269–283. doi: 10.1521/pedi.2008.22.3.269 Shields, A., & Cicchetti, D. (1998). Reactive aggression among maltreated children: The contributions of attention and emotion dysregulation. Journal of Clinical Child Psychology, 27(4), 381–395. doi: 10.1207/s15374424jccp2704_2 Shoulberg, E. K., Sijtsema, J. J., & Murray-Close, D. (2011). The association between valuing popularity and relational aggression: The moderating effects of actual popularity and physiological reactivity to exclusion. Journal of Experimental Child Psychology, 110, 20–37. doi: 10.1016/j.jecp.2011.03.008 Shuster, M. M., Li, Y., & Shi, J. (2012). Maternal cultural values and parenting practices: Longitudinal associations with Chinese adolescents' aggression. Journal of Adolescence, 35(2), 345–355. doi: 10.1016/j.adolescence.2011.08.006 Sijtsema, J. J., Ojanen, T., Veenstra, R., Lindenberg, S., Hawley, P. H., & Little, T. D. (2010). Forms and functions of aggression in adolescent friendship selection and influence: A longitudinal social network analysis. Social Development, 19(3), 515–534. doi: 10.1111/j.1467–9507.2009.00566.x Sijtsema, J. J., Shoulberg, E. K., & Murray-Close, D. (2011). Physiological reactivity and different forms of aggression in girls: Moderating roles of rejection sensitivity and peer rejection. Biological Psychology, 86(3), 181–192. doi: 10.1016/j.biopsycho.2010.11.007 Sijtsema, J. J., Veenstra, R., Lindenberg, S., van Roon, A. M., Verhulst, F. C., Ormel, J., & Riese, H. (2010). Mediation of sensation seeking and behavioral inhibition on the relationship between heart rate and antisocial behavior: The TRAILS study. Journal of the American Academy of Child & Adolescent Psychiatry, 49(5), 493–502. doi: 10.1016/j.jaae.2010.02.005 Silverthorn, P., & Frick, P. J. (1999). Developmental pathways to antisocial behavior: The delayed-onset pathway in girls. Development and Psychopathology, 11, 101–126. doi: 10.1017/S0954579499001972 Simon, V. A., Aikins, J. W., & Prinstein, M. J. (2008). Romantic partner selection and socialization during early adolescence. Child Development, 79(6), 1676–1692. doi: 10.1111/j.1467–8624.2008.01218.x Skara, S., Pokhrel, P., Weiner, M. D., Sun, P., Dent, C. W., & Sussman, S. (2008). Physical and relational aggression as predictors of drug use: Gender differences among high school students. Addictive Behaviors, 33(12), 1507–1515. doi: 10.1016/j.addbeh.2008.05.014 Slaby, R. G., & Guerra, N. G. (1988). Cognitive mediators of aggression in adolescent offenders: I. Assessment. Developmental Psychology, 24(4), 580–588. doi: 10.1037/0012– 1649.24.4.580

Smith, S. L., & Donnerstein, E. (1998). Harmful effects of exposure to media violence: Learning of aggression, emotional desensitization, and fear. In R. G. Geen & E. Donnerstein (Eds.), Human aggression: Theories, research, and implications for social policy (pp. 167– 202). San Diego, CA: Academic. Smith, R. L., Rose, A. J., & Schwartz-Mette, R. A. (2010). Relational and overt aggression in childhood and adolescence: Clarifying mean-level gender differences and associations with peer acceptance. Social Development, 19, 243–269. doi: 10.1111/j.1467–9507.2009.00541.x Smits, I., Doumen, S., Luyckx, K., Duriez, B., & Goossens, L. (2011). Identity styles and interpersonal behavior in emerging adulthood: The intervening role of empathy. Social Development, 20(4), 664–684. doi: 10.1111/j.1467–9507.2010.00595.x Soensens, B., Vansteenkiste, M., Goossens, L., Duriez, B., & Niemiec, C. P. (2008). The intervening role of relational aggression between psychological control and friendship quality. Social Development, 17(3), 661–681. doi: 10.1111/j.1467–9507.2007.00454.x Spieker, S. J., Campbell, S. B., Vandergrift, N., Pierce, K. M., Cauffman, E., Susman, E. J., & Roisman, G. I. (2012). Relational aggression in middle childhood: Predictors and adolescent outcomes. Social Development, 21(2), 354–375. doi: 10.1111/j.1467–9507.2011.00631.x Sroufe, L. A. (1997). Psychopathology as an outcome of development. Development and Psychopathology, 9(2), 251–268. doi: 10.1017/S0954579497002046 Sroufe, L. A., Egeland, B., & Carlson, E. (1999). One social world: The integrated development of parent-child and peer relationships. In W. A. Collins & B. Laursen (Eds.), Relationships as developmental context: The 30th Minnesota symposium on child psychology (pp. 241–262). Mahwah, NJ: Erlbaum. Stauffacher, K., & DeHart, G. B. (2005). Preschoolers' relational aggression with siblings and with friends. Early Education and Development, 16(2), 185–206. doi: 10.1207/s15566935eed1602_5 Stauffacher, K., & DeHart, G. B. (2006). Crossing social contexts: Relational aggression between siblings and friends during early and middle childhood. Journal of Applied Developmental Psychology, 27(3), 228–240. doi: 10.1016/j.appdev.2006.02.004 Stepp, S. D., Pilkonis, P. A., Hipwell, A. E., Loeber, R., & Stouthamer-Loeber, M. (2010). Stability of borderline personality disorder features in girls. Journal of Personality Disorders, 24(4), 460–472. doi: 10.1521/pedi.2010.24.4.460 Stickle, T. R., Marini, V. A., & Thomas, J. N. (2012). Gender differences in psychopathic traits, types, and correlates of aggression among adjudicated youth. Journal of Abnormal Child Psychology, 40, 513–525. doi: 10.1007/s10802–011–9588–1 Storch, E. A., Bagner, D. M., Geffken, G. R., & Baumeister, A. L. (2004). Association between overt and relational aggression and psychological adjustment in undergraduate college

students. Violence and Victims, 19(6), 689–700. doi: 10.1891/vivi.19.6.689.66342 Strayer, J., & Strayer, F. F. (1978). Social aggression and power relations among preschool children. Aggressive Behavior, 4(2), 173–182. doi: 10.1002/1098–2337(1978)4:2 Sugimura, N., & Rudolph, K. D. (2012). Temperamental differences in children's reactions to peer victimization. Journal of Clinical Child and Adolescent Psychology, 41(3), 314–328. doi: 10.1080/15374416.2012.656555 Sullivan, T. N., Farrell, A. D., & Kliewer, W. (2006). Peer victimization in early adolescence: Association between physical and relational victimization and drug use, aggression, and delinquent behaviors among urban middle school students. Development and Psychopathology, 18, 119–137. doi: 10.1017/S095457940606007X Susman, E. J., Dockray, S., Schiefelbein, V. L., Herwehe, S., Heaton, J. A., & Dorn, L. D. (2007). Morningness/eveningness, morning-to-afternoon cortisol ratio, and antisocial behavior problems during puberty. Developmental Psychology, 43(4), 811–822. doi: 10.1037/0012– 1649.43.4.811 Sutton, J., Smith, P. K., & Swettenham, J. (1999). Social cognition and bullying: Social inadequacy or skilled manipulation? British Journal of Developmental Psychology, 17(3), 435–450. doi: 10.1348/026151099165384 Tackett, J. L., Daoud, S. L. S. B., De Bolle, M., & Burt, S. A. (2013). Is relational aggression part of the externalizing spectrum? A bifactor model of youth antisocial behavior. Aggressive Behavior, 39(2), 149–159. doi: 10.1002/ab.21466 Tackett, J. L., Kushner, S. C., Herzhoff, K., Smack, A. J., & Reardon, K. W. (2014). Viewing relational aggression through multiple lenses: Temperament, personality, and personality pathology. Development and Psychopathology, 26, 863–877. doi: 10.1017/S0954579414000443 Tackett, J. L., & Ostrov, J. M. (2010). Measuring relational aggression in middle childhood in a multi-informant multi-method study. Journal of Psychopathology and Behavioral Assessment, 32(4), 490–500. doi: 10.1007/s10862–010–9184–7 Tackett, J. L., Waldman, I. D., & Lahey, B. B. (2009). Etiology and measurement of relational aggression: A multi-informant behavior genetic investigation. Journal of Abnormal Psychology, 118(4), 722–733. doi: 10.1037/a0016949 Terranova, A. M., Morris, A. S., & Boxer, P. (2008). Fear reactivity and effortful control in overt and relational bullying: A six-month longitudinal study. Aggressive Behavior, 34, 104– 115. doi: 10.1002/ab.20232 Tiet, Q. Q., Wasserman, G. A., Loeber, R., McReynolds, L. S., & Miller, L. S. (2001). Developmental and sex differences in types of conduct problems. Journal of Child & Family Studies, 10(2), 181–197. doi: 10.1023/A:1016637702525

Tomada, G., & Schneider, B. H. (1997). Relational aggression, gender, and peer acceptance: Invariance across culture, stability over time, and concordance among informants. Developmental Psychology, 33(4), 601–609. doi: 10.1037/0012–1649.33.4.601 Trompetter, H., Scholte, R., & Westerhof, G. (2011). Resident-to-resident relational aggression and subjective well-being in assisted living facilities. Aging & Mental Health, 15, 59–67. doi: 10.1080/13607863.2010.501059 Tseng, W.-L., Banny, A. M., Kawabata, Y., Crick, N. R., & Gau, S. S.-F. (2013). A crosslagged structural equation model of relational aggression, physical aggression, and peer status in a Chinese culture. Aggressive Behavior, 39(4), 301–315. doi: 10.1002/ab.21480 Tseng, W.-L., Kawabata, Y., Gau, S. S.-F., Banny, A. M., Lingras, K. A., & Crick, N. R. (2012). Relations of inattention and hyperactivity/impulsivity to preadolescent peer functioning: The mediating roles of aggressive and prosocial behaviors. Journal of Clinical Child and Adolescent Psychology, 41(3), 275–287. doi: 10.1080/15374416.2012.656556 Underwood, M. K. (2003). The comity of modest manipulation, the importance of distinguishing among bad behaviors. Merrill-Palmer Quarterly, 49(3), 373–389. doi: 10.1353/mpq.2003.0016 Underwood, M. K. (2004). III. Glares of contempt, eye rolls of disgust, and turning away to exclude: Non-verbal forms of social aggression among girls. Feminism and Psychology, 14(3), 371–375. doi: 10.1177/0959353504044637 Underwood, M. K., Beron, K. J., Gentsch, J. K., Galperin, M. B., & Risser, S. D. (2008). Family correlates of children's social and physical aggression with peers: Negative interparental conflict strategies and parenting styles. International Journal of Behavioral Development, 32(6), 549–562. doi: 10.1177/0165025408097134 Underwood, M. K., Beron, K. J., & Rosen, L. H. (2009). Continuity and change in social and physical aggression from middle childhood through early adolescence. Aggressive Behavior, 35(5), 357–375. doi: 10.1002/ab.20313 Underwood, M. K., Beron, K. J., & Rosen, L. H. (2011). Joint trajectories for social and physical aggression as predictors of adolescent maladjustment: Internalizing symptoms, rulebreaking behaviors, and borderline and narcissistic personality features. Development and Psychopathology, 23(2), 659–678. doi: 10.1017/S095457941100023X Underwood, M. K., Galen, B. R., & Paquette, J. A. (2001). Top ten challenges for understanding gender and aggression in children: Why can't we all just get along? Social Development, 10(2), 248–266. doi: 10.1111/1467–9507.00162 Updegraff, K. A., Thayer, S. M., Whiteman, S. D., Denning, D. J., & McHale, S. M. (2005). Relational aggression in adolescents' sibling relationships: Links to parent-adolescent relationship quality. Family Relations 54(3), 373–385. doi: 10.1111/j.1741–

3729.2005.00324.x Vaillancourt, T., Brendgen, M., Boivin, M., & Tremblay, R. E. (2003). A longitudinal confirmatory factor analysis of indirect and physical aggression: Evidence of two factors over time? Child Development, 74, 1628–1638. doi: 10.1046/j.1467–8624.2003.00628.x Vaillancourt, T., Miller, J. L., Fagbemi, J., Côté, S., & Tremblay, R. E. (2007). Trajectories and predictors of indirect aggression: Results from a nationally representative longitudinal study of Canadian children aged 2–10. Aggressive Behavior, 33(4), 314–326. doi: 10.1002/ab.20202 Vaillancourt, T., & Sunderani, S. (2011). Psychopathy and indirect aggression: The roles of cortisol, sex, and type of psychopathy. Brain and Cognition, 77(2), 170–175. doi: 10.1016/j.bandc.2011.06.009 van Geert, P., & Steenbeek, H. (2005). Explaining after by before: Basic aspects of a dynamic systems approach to the study of development. Developmental Review, 25(3–4), 408–442. doi: 10.1016/j.dr.2005.10.003 Velásquez, A. M., Santo, J. B., Saldarriaga, L. M., López, L. S., & Bukowksi, W. M. (2010). Context-dependent victimization and aggression: Differences between all-girl and mixed-sex schools. Merrill-Palmer Quarterly, 56(3), 283–302. doi: 10.1353/mpq.0.0054 Verona, E., & Kilmer, A. (2007). Stress exposure and affective modulation of aggressive behavior in men and women. Journal of Abnormal Psychology, 116(2), 410–421. doi: 10.1037/0021–843X.116.2.410 Vitaro, F., Gendreau, P. L., Tremblay, R. E., & Oligny, P. (1998). Reactive and proactive aggression differentially predict later conduct problems. Journal of Child Psychology and Psychiatry, 39(3), 377–385. doi: 10.1017/S0021963097002102 Waasdorp, T. E., Baker, C. N., Paskewich, B. S., & Leff, S. S. (2013). The association between forms of aggression, leadership, and social status among urban youth. Journal of Youth and Adolescence, 42, 263–274. doi: 10.1007/s10964–012–9837–9 Waasdorp, T. E., & Bradshaw, C. P. (2011). Examining student responses to frequent bullying: A latent class approach. Journal of Educational Psychology, 103(2), 336–352. doi: 10.1037/a0022747 Walcott, C. M., Upton, A., Bolen, L. M., & Brown, M. B. (2008). Associations between peerperceived status and aggression in young adolescents. Psychology in the Schools, 45(6), 550– 561. doi: 10.1002/pits.20323 Warren, G. C., & Clarbour, J. (2009). Relationship between psychopathy and indirect aggression use in a noncriminal population. Aggressive Behavior, 35(5), 408–421. doi: 10.1002/ab.20317

Werner, N. E., Bumpus, M. F., & Rock, D. (2010). Involvement in Internet aggression during early adolescence. Journal of Youth and Adolescence, 39(6), 607–619. doi: 10.1007/s10964– 009–9419–7 Werner, N. E., & Crick, N. R. (1999). Relational aggression and social-psychological adjustment in a college sample. Journal of Abnormal Psychology, 108(4), 615–623. doi: 10.1037/0021–843X.108.4.615 Werner, N. E., & Crick, N. R. (2004). Maladaptive peer relationships and the development of relational and physical aggression during middle childhood. Social Development, 13, 495– 514. doi: 10.1111/j.1467–9507.2004.00280.x Werner, N. E., & Grant, S. (2009). Mothers' cognitions about relational aggression: Associations with discipline responses, children's normative beliefs, and peer competence. Social Development, 18, 77–98. doi: 10.1111/j.1467–9507.2008.00482.x Werner, N. E., & Hill, L. G. (2010). Individual and peer group normative beliefs about relational aggression. Child Development, 81(3), 826–836. doi: 10.1111/j.1467– 8624.2010.01436.x Werner, N. E., Senich, S., & Przepyszny, K. A. (2006). Mothers' responses to preschoolers' relational and physical aggression. Journal of Applied Developmental Psychology, 27(3), 193–208. doi: 10.1016/j.appdev.2006.02.002 White, D. D., Gallup, A. C., & Gallup, G. G. (2010). Indirect peer aggression in adolescence and reproductive behavior. Evolutionary Psychology, 8, 49–65. Williams, S. C., Lochman, J. E., Phillips, N. C., & Barry, T. D. (2003). Aggressive and nonaggressive boys' physiological and cognitive processes in response to peer provocations. Journal of Clinical Child & Adolescent Psychology, 32(4), 568–576. doi: 10.1207/S15374424JCCP3204_9 Wright, M. F., & Li, Y. (2013). The association between cyber victimization and subsequent cyber aggression: The moderating effect of peer rejection. Journal of Youth and Adolescence, 42(5), 662–674. doi: 10.1007/s10964–012–9903–3 Woods, T., Coyle, K., Hoglund, W., & Leadbeater, B. J. (2007). Changing the contexts of peer victimization: The effects of a primary prevention program on school and classroom levels of victimization. In J. E. Zins, M. J. Elias, & C. A. Maher (Eds.), Bullying, victimization, and peer harassment: A handbook of prevention and intervention (pp. 369–388). New York, NY: Haworth Press. Yeung, R. S., & Leadbeater, B. J. (2007). Does hostile attributional bias for relational provocations mediate the short-term association between relational victimization and aggression in preadolescence? Journal of Youth and Adolescence, 36(8), 973–983. doi: 10.1007/s10964–006–9162–2

Zahn-Waxler, C., Crick, N. R., Shirtcliff, E. A., & Woods, K. E. (2006). The origins and development of psychopathology in females and males. In D. Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology: Theory and method (Vol. 1, 2nd ed., pp. 76–138). Hoboken, NJ: Wiley. Zalecki, C. A., & Hinshaw, S. P. (2004). Overt and relational aggression in girls with attention deficit hyperactivity disorder. Journal of Clinical Child & Adolescent Psychology, 33, 125– 137. doi: 10.1207/S15374424JCCP3301_12 Zimmer-Gembeck, M. J., & Duffy, A. L. (2014). Heightened emotional sensitivity intensifies associations between relational aggression and victimization among girls but not boys: A longitudinal study. Development and Psychopathology, 26, 661–673. doi: 10.1017/S0954579414000303 Zimmer-Gembeck, M. J., Geiger, T. C., & Crick, N. R. (2005). Relational and physical aggression, prosocial behavior, and peer relations: Gender moderation and bidirectional associations. Journal of Early Adolescence, 25, 421–452. doi: 10.1177/0272431605279841 Zwaan, M., Dijkstra, J. K., & Veenstra, R. (2013). Status hierarchy, attractiveness hierarchy and sex ratio: Three contextual factors explaining the status–aggression link among adolescents. International Journal of Behavioral Development, 37(3), 211–221. doi: 10.1177/0165025412471018

Chapter 14 Culture, Peer Relationships, and Developmental Psychopathology Xinyin Chen and Cindy H. Liu The writing of this chapter was supported by grants from the National Science Foundation (#BCS-1225620) and the Social Sciences and Humanities Research Council of Canada (SSHRC) to X. Chen. THEORETICAL PERSPECTIVES ON CULTURE, PEER RELATIONSHIPS, AND ADAPTIVE AND MALADAPTIVE DEVELOPMENT The Role of Culture in Development: Traditional Perspectives Culture, Peer Relationships, and Adaptive and Maladaptive Development: The Contextual-Developmental Perspective SOCIOEMOTIONAL FUNCTIONING AND PROBLEMS IN PEER SETTINGS ACROSS CULTURES Shyness-Inhibition, Social Anxiety, and Loneliness Aggressive, Violent, and Other Externalizing Behaviors Peer Conflict and Conflict Resolution Bullying and Victimization Peer Involvement, Isolation, and Friendship Exclusivity CULTURE AND PARENTAL ATTITUDES Parental Attitudes Toward Shyness-Inhibition and Social Anxiety Parental Attitudes Toward Aggression, Anger, and Self-Regulation Parental Attitudes Toward Peer Conflict Parental Attitudes Toward Bullying and Victimization Parental Attitudes Toward Peer Involvement CULTURE, PEER EVALUATION, AND THE REGULATORY FUNCTION OF PEER INTERACTIONS Peer Evaluations The Regulatory Function of Culturally Directed Peer Interactions CHILDREN'S PEER EXPERIENCES AND ADAPTIVE AND MALADAPTIVE OUTCOMES: THE ROLE OF CULTURE

Shyness-Inhibition, Unsociability, and Adjustment Aggression, Self-Control, and Adjustment Peer Conflict and Adjustment Bullying and Victimization and Adjustment Peer Group Involvement, Friendships, and Adjustment Peer Relationships and Resilience CONCLUSIONS, PRACTICAL IMPLICATIONS, AND FUTURE DIRECTIONS REFERENCES The role of culture in human development is one of the major issues in sociology, anthropology, psychology, and clinical science. Research findings have indicated considerable variations in children's and adolescents' behaviors, emotions, and cognitions across cultures. Cultural norms and values are involved in development through various processes, such as facilitation and suppression of specific behaviors (Weisz, Weiss, Suwanlert, & Chaiyasit, 2006). Culture also provides a frame of reference for social judgments of behaviors and thus ascribes meanings to the behaviors (X. Chen & French, 2008). Therefore, whether a behavior is viewed as adaptive or maladaptive depends on cultural context (e.g., García Coll, Akerman, & Cicchetti, 2000; Kleinman, 1980). Moreover, how adaptive and maladaptive behaviors develop in a society or community is determined, in part, by cultural factors. Developmental theorists and researchers have traditionally been interested in the socialization role of adults, especially parents and educators, in transmitting cultural values to the young generation (e.g., Goodnow, 1997; LeVine et al., 1994; Vygotsky, 1978). However, the processes of cultural influence on development, particularly in socioemotional areas, are more complicated and more comprehensive than the transmission of the cultural system from senior members of the society. Many cultural norms and values, such as self-control, that serve to promote positive interpersonal interactions and group functioning do not have inherent benefit per se and thus may not be readily appreciated by children. Moreover, adult influence becomes more distal as children develop greater independence with age and as children engage in more activities outside the home and classroom. Thus, cultural norms and values that are endorsed in peer groups, which may or may not be compatible with the adults' cultural system, become increasingly important in guiding children's and adolescents' behaviors. X. Chen (2012) and X. Chen and colleagues (e.g., X. Chen & French, 2008) argued that cultural influence on individual development involves peer interactions and relationships. According to this view, while peer interactions and relationships contribute to children's and adolescents' adaptive and maladaptive development, culture affects the processes of peer interactions and features of peer relationships. The significance of peer interactions and relationships for social, behavioral, and cognitive development has long been well recognized in the literature (e.g., Hartup, 1992; Piaget, 1932; Rubin, Bukowski, & Parker, 2006). Peer interactions provide opportunities for children to

learn social and problem-solving skills, such as negotiation and cooperation, from one another (e.g., Hartup, 1992). The experience of peer interactions helps children understand standards for appropriate behaviors in different settings. Moreover, social affiliations established through interaction are a main source of feelings of security and belonging, which in turn constitute a basis for the development of psychological well-being (e.g., Sullivan, 1953). Thus, achieving and maintaining appropriate behaviors and establishing desirable relationships in the peer context are major developmental tasks in childhood and adolescence (Masten & Tellegen, 2012; Rubin, Bukowski, & Bowker, 2015). Failure to behave competently in peer interactions, obtain acceptance in the peer group, or form constructive relationships and networks is a significant indication of developmental psychopathology and may have enduring and cascading effects on children's adjustment in broad areas (Masten & Cicchetti, 2010; Parker, Rubin, Erath, Wojslawowicz, & Buskirk, 2006). Consistent with these arguments, empirical results have indicated that peer interactions and relationships are associated with various developmental outcomes (Coplan, Prakash, O'Neil, & Armer, 2004; Dodge, Greenberg, & Malone, 2008; Rubin, Bukowski, et al., 2006; Ladd & Troop-Gordon, 2003). Peer interactions and relationships are shaped, however, by cultural conventions, norms, and values in the society (X. Chen & French, 2008; Edwards, de Guzman, Brown, & Kumru, 2006). According to Hinde (1987), different levels of social experiences are embedded within the cultural system. Similarly, Schneider, Smith, Poisson, and Kwan (1997) pointed out that culture may affect many aspects of peer interactions and relationships. Indeed, cross-cultural research has indicated that cultural beliefs and practices, particularly those pertaining to socialization and developmental goals, affect the nature, function, and features of peer interactions and relationships such as friendships and social networks in children and adolescents (e.g., DeRosier & Kupersmidt, 1991; Farver, Kim, & Lee, 1995; French, Pidada, & Victor, 2005). For example, cultural norms and values may serve as a basis for the interpretation of particular behaviors and for the judgment of the appropriateness of these behaviors in peer interactions. The social interpretation and evaluation in turn regulate the processes of peer interactions and the formation of dyadic and group relationships. In addition, the role of culture is reflected in how it affects the relations between peer experiences and adjustment outcomes. In this chapter, we focus on how culture is involved in children's peer interactions and relationships and the implications of peer interactions and relationships for the development of social, behavioral, and psychological problems. We first discuss some conceptual issues in the study of culture, peer relationships, and developmental psychopathology. Then we review research on the prevalence of major social behaviors and problems in peer interactions and relationships among children across cultures. Next, we discuss culturally directed social attitudes and responses (e.g., parental attitudes, peer evaluations and regulations) toward specific behaviors within the peer context. In our discussion, we pay particular attention to how macro-level social and cultural conditions and their changes, through directing the social interaction processes, determine the functional meanings of behaviors and promote or undermine their development. Then we focus on the developmental outcomes of peer experiences in different cultures. The chapter concludes with a discussion of general issues,

implications, and future directions in the study of culture, peer relationships, and developmental psychopathology.

Theoretical Perspectives on Culture, Peer Relationships, and Adaptive and Maladaptive Development Two major theories of culture and human development are the socioecological theory (Bronfenbrenner & Morris, 2006) and the sociocultural theory (M. Cole, 1996; Rogoff, 2003; Vygotsky, 1978). According to both theories, culture affects socialization beliefs and practices, which in turn contribute to developmental outcomes in various areas. In addition to these broadband theories, a cultural anthropological perspective (Benedict, 1934) focuses on how culture affects the judgment of behaviors in different societies. These theoretical perspectives have guided research on culture and normal and abnormal development for the past 50 years. X. Chen (2012) and X. Chen and French (2008) recently proposed a contextual-developmental perspective focusing on peer interaction as a context for cultural influence on socioemotional development. This perspective emphasizes the role of the social evaluation and regulation processes in peer interaction in facilitating the links between culture and development.

The Role of Culture in Development: Traditional Perspectives According to the socioecological theory (Bronfenbrenner & Morris, 2006), the cultural beliefs and practices that are endorsed within a society or community, as a part of the socioecological environment, play an important role in children's socioemotional and cognitive development. In the early version of the theory (e.g., Bronfenbrenner, 1979), culture was considered a distal influence in the outmost layer of the environment in which the child did not directly participate. In more recent conceptualizations, however, cultural factors have been integrated with proximal socialization forces such as the community and family (Tietjen, 2006). Culture influences human development through organizing various social settings and activities. From the socioecological perspective, providing desirable cultural conditions is critical for promoting adaptive development and for preventing developmental problems, although it is unclear what specific cultural aspects or processes may lead to normal and abnormal outcomes. Largely consistent with the socioecological perspective, Super and Harkness (1986) proposed a developmental niche model, which focuses on three main interacting subsystems: the physical and social settings; the historically constituted customs and practices of child care and child rearing; and the psychology of the caretakers, particularly parental ethnotheories shared with the community. Culture may affect human development through these subsystems. According to Super and Harkness, deprived or disadvantaged settings, disorganized childrearing practices, and inappropriate socialization beliefs are the main risk factors in psychopathological development. The sociocultural theory focuses on the internalization of cultural systems, such as language and symbols, from the interpersonal level to the intrapersonal level (Vygotsky, 1978).

Participation in social and cultural practices may exert a significant influence on the process and pattern of human development. Changes in sociocultural structures and associated social practices may lead to the reorganization of mental systems and the formation of new psychological functions (Luria, 1976). For example, the transformation of traditional rural villages to communities with more complex and planned social activities may produce new forms of knowledge and ways of thinking. Consistent with the sociocultural perspective, crosscultural studies (e.g., Beach, 1995; Luria, 1976; Rogoff, 2003; Scribner & Cole, 1981) have shown that, relative to their counterparts who participated in formal education or urbanized activities, children and adults in villages in Africa, Asia, and Central and South America who lived traditional lifestyles tended to display more concrete practical or graphic-functional thinking and context-specific calculation or visual representation that were constrained by the physical features of the circumstances. The mental processes of individuals in these societies became more decontextualized, sophisticated, and logical as they engaged in more commercial activities and formal school learning (Luria, 1976; Vygotsky, 1978). A long-held perspective in psychopathology, proposed mainly by Benedict (1934) and Mead (1928), stresses cultural relativity in the judgment of normal and abnormal behaviors. According to this perspective, normality is defined by culture because a behavior viewed as abnormal in one culture may be viewed as normal in another. In all societies, there are substantial individual variations on behaviors and personality characteristics. However, different societies may place different values on specific behaviors and characteristics and define specific ranges of these behaviors and characteristics as normal or desired. Normality represents a segment of human functioning that is approved within a culture, and abnormality is the term for the segment that is incompatible with the norm in the society. Thus, normal behaviors and characteristics are the ones that are considered within the limits of the society; those beyond the limits are viewed as problematic. Accordingly, individuals who display culturally acceptable and desirable behaviors are likely to receive social approval and prestige and to obtain social status within the society or the community. In contrast, individuals who display behaviors not in agreement with the ones selected in that society or community are regarded as abnormal, even though these behaviors are valued in other cultures. As indicated by Benedict (1973), “We must face the fact that even our normality is man-made, and is of our own seeking” (p. 76). During the process of socialization, culture serves to facilitate the development of valued behaviors and suppress those behaviors considered abnormal and unacceptable. Benedict argued that most individuals are plastic to the molding force of the society into which they are born. Consequently, the vast majority of the individuals in any group are socialized to fit with the cultural expectations (p. 74). Benedict's relativist perspective is not necessarily incompatible with the universalist view that mental or psychological disorders have universal core symptoms although cultures may vary on the presentations and perceptions of the disorders. The universalist view asserts that the underlying psychopathological problems are the same, but the manifestation of the problems and the threshold of what is judged as problematic may differ across cultures (e.g., Bird, 2002; Roberts & Roberts, 2007). For example, it has been reported that the rates of attentiondeficit/hyperactivity disorder in Hong Kong are double those reported in some Western

countries (T. Ho, Leung, Luk, & Taylor, 1996). According to both Benedict's (1934) and the universalist views, the different rates may be due to the threshold effect. Since Asian cultures value the suppression of emotions such as anger and disruptive behaviors, parents and mental health practitioners may have a lower threshold for the judgment of hyperactive behavior, which increases the likelihood of reporting hyperactive behaviors (Canino & Alegría, 2008). Indeed, compared to American clinicians, when they judged the same vignettes, Chinese and Indonesian clinicians gave significantly higher scores for hyperactive behavior (Mann, Ideda, Mueller, & Takahashi, 1992). A more thorough relativist view (e.g., Kleinman & Kleinman, 1991; Weisz et al., 2006) claims that culture affects the manifestation and judgment of a specific behavior as well as the development of the behavior per se. Cultural conditions shape the occurrence, magnitude, duration, and form of the behavior and social responses to the behavior. Thus, a symptom or syndrome may be observed in some cultures but not in others. Social and cultural factors can be important determinants of behavioral and psychological problems. Between the universalist and relativist views, many researchers and practitioners believe that whereas problems derived from neural pathology (e.g. autism, schizophrenia) may be universal, more common problems are likely to be affected by social and cultural circumstances (García Coll et al., 2000; Rutter & Nikapota, 2002; also see Canino & Alegría, 2008). How a specific behavior or problem occurs and is viewed in a society and how culture plays a role may depend on the nature of the behavior or problem. Our discussion of culture and developmental psychopathology in the next sections focuses mainly on relatively common behaviors, such as aggression, defiance, shyness, and social anxiety, although there is evidence indicating biological foundations for the development of these behaviors (e.g., Fox, Henderson, Marshall, Nichols, & Ghera, 2005). Researchers in the field of developmental psychopathology have recently noticed that distinct functions observed in cultural and ethnic minority populations, such as Asian and Latino children in the United States, may represent strengths, rather than deficits as traditionally viewed, in their adaptation to the environment (e.g., García Coll et al., 2000; X. Chen & Tse, 2008). Thus, largely consistent with the relativist perspective, it has been argued that psychopathological issues, such as risk, mental illness, resilience, and protective factors, should be examined with particular cultural standards for minority groups (García Coll et al., 2000). According to Cicchetti and Toth (2009), to fully comprehend the complexities in the experiences of psychopathology, it is necessary to investigate biological, psychological, and social aspects of normal and abnormal development at multiple levels of analysis. The analysis at the cultural level may provide useful information about how different aspects of development are associated with, and facilitate and constrain, each other in specific contexts. Research on culture and developmental psychopathology not only helps us understand the broad, dynamic processes underlying individual adjustment and maladjustment but also informs the diagnosis and treatment of problems.

Culture, Peer Relationships, and Adaptive and Maladaptive Development: The Contextual-Developmental Perspective

The contextual-developmental perspective, developed by X. Chen (2012) and X. Chen and French (2008) and based mainly on the socioecological and sociocultural theories (Bronfenbrenner & Morris, 2006; Vygotsky, 1978), emphasizes the links between cultural values and children's socioemotional functioning and the role of the social interactions in mediating the links. This perspective focuses on two fundamental dimensions of socioemotional functioning: social initiative and self-control. Social initiative is concerned with the tendency to spontaneously initiate and maintain social participation, especially in stressful settings. Whereas some children may readily engage in interactions in potentially challenging situations, others may experience internal anxiety and fear, leading to exhibition of low levels of social initiative (Asendorpf, 1990). The display of shy-inhibited and withdrawn behaviors is a major indication of low social initiative. Self-control represents the regulatory ability to modulate behavioral and emotional reactivity to maintain appropriate behavior in social activities. The dimension of control, indicated often by the display of compliantcooperative and defiant-aggressive behaviors, is concerned with fitting in with others and, more broadly, achieving interpersonal harmony and group well-being. Different cultures may value social initiative and norm-based behavioral control in children and adolescents to different extents. In Western self-oriented or individualistic cultures, where acquiring autonomy and assertive skills is an important socialization goal, social initiative is viewed as a major index of competence; the display of inhibited or restrained behavior is often considered immature and incompetent (X. Chen et al., 1998). Self-regulation and control, in general, are perceived as necessary for adaptive functioning. However, individuals are encouraged to learn how to maintain a balance between the needs of the self and those of others. Consequently, behavioral control is less appreciated and endorsed, especially when it conflicts with the attainment of individual goals (Triandis, 1990). In group-oriented cultures, social initiative may not be highly emphasized because it may not have evident effects on the harmony of the group. To maintain group harmony, however, individuals need to restrain personal desires to address the interests of others, and thus self-control is more strongly emphasized (X. Chen, Rubin, et al., 2003; D. Y. E. Ho, 1986). The lack of control is viewed as highly unacceptable. Cultural values of social initiative and self-control may affect adults' and children's attitudes toward specific aspects of socioemotional functioning, including aggression-defiance (high social initiative and low control), shyness-inhibition (low social initiative and adequate control to constrain behavioral and emotional reactivity to self), sociable-cooperative behaviors (active social initiative with effective control), and affective disturbances such as depression and loneliness (low social initiative and low control). According to the contextual-developmental perspective, social evaluation and response processes play a significant role in maintaining the links between culture and human development. During interactions, adults and peers evaluate children's behaviors according to cultural norms and values that are endorsed in the community. Moreover, adults and peers across cultures may respond differently to specific behaviors and express different attitudes toward children who display these behaviors. To acquire acceptance and approval in the group, children need to understand social standards and expectations and to adjust their behaviors according to the standards and expectations. Thus, social evaluations and responses,

particularly in peer context, serve to regulate children's behaviors and their development (X. Chen, 2012). Whether and to what extent children can regulate their behaviors or behavioral styles according to culturally directed social evaluations is associated with adjustment outcomes. Children whose behaviors are incompatible with group expectations are likely to receive negative feedback from others, which creates a pressure on children to alter their behaviors. In antisocial or other socially deviant groups, the social interaction processes may promote individual deviant behaviors, which may lead to problems in the larger social environment, especially in the long run. Relative to other members in the group, children who are able to resist the peer pressure may be more resilient to maladaptive group socialization. In prosocial groups, however, peer evaluations and responses in interactions serve to facilitate the development of socially constructive behaviors and characteristics. Children who can regulate their behaviors to improve their social relationships with others may increase their opportunities to learn from others and gain emotional support and other benefits, which reduce the likelihood of developing social, cognitive, and psychological problems. However, children who fail to do so may obtain increasingly unfavorable experience in interactions, which may elicit frustration, anger, and other negative emotions. These negative emotional reactions may lead to further externalizing problems, such as aggression, if directed toward others; internalizing problems, such as loneliness and depression, if directed toward the self, or both. By focusing on peer interaction processes, the contextual-developmental perspective emphasizes the active role of children in socialization and development. The active role of children may be reflected in their sensitivity and response to social influence. With age, the active role of children becomes increasingly evident through their participation in adopting existing cultures and constructing new cultures for social interactions (Corsaro & Nelson, 2003). As argued by Tamis-LeMonda et al. (2008), the construction of new cultures may be most likely to take place in social activities of children with different backgrounds. Migration and communication across nations during globalization have created a changing context with diverse values for children and adolescents in most societies. As a result, exposure to different beliefs and lifestyles has become a part of the youth experience. The cultural systems that youth develop through incorporating different, and perhaps complementary, values and behavioral norms provide guidance for them to engage in mutual evaluations in interaction and to maintain their behaviors according to cultural norms.

Socioemotional Functioning and Problems in Peer Settings Across Cultures In general, behaviors that are formed on the basis of high social initiative and high self-control, such as cooperative and prosocial behavior, are regarded as appropriate and desirable in most societies (e.g., Eisenberg, Fabes, & Spinrad, 2006), although the extent to which they are valued may vary across cultures. Behaviors formed on the basis of low levels of social initiative, self-control, or both and categorized as internalizing and externalizing are typically

viewed as maladaptive and abnormal in Western cultures (Achenbach, 2008; Cicchetti, 2006) but not necessarily in other cultures. The cultural values of these behaviors play a major role in determining how they are exhibited, particularly in social situations. A number of studies have been conducted to examine cross-cultural similarities and differences in the display of children's behaviors and problems. Most of the studies, however, have relied on parental ratings or self-reports, which suffer from methodological problems, such as judgmental biases, response style biases, and reference group biases (e.g., Schneider, French, & Chen, 2006). The methodological weaknesses and problems have led to inconsistencies and confusions of the results. Nevertheless, some patterns have emerged from cross-cultural studies of children in Asian, Latino, European, and North American societies. For example, studies based on parental reports (e.g., Gartstein et al., 2006) indicated that children in some Asian cultures were less emotionally expressive than children in European and North American cultures in early childhood. These results are consistent with what has been found in observational studies conducted by Camras and her colleagues (Camras et al., 1998; Camras, Chen, Bakeman, Norris, & Cain, 2006) in East Asian and European American infants. The cross-cultural differences in the early years are interesting because they may indicate the developmental origins of socioemotional functioning and problems among children in specific societies.

Shyness-Inhibition, Social Anxiety, and Loneliness Researchers have observed that children display different reactions to stressful or challenging situations (e.g., Asendorpf, 1991; Fox et al., 2005; Kagan, 1997; Stevenson-Hinde, Shouldice, & Chicot, 2011). Whereas some children are relaxed and display relatively little distress, others tend to be vigilant, anxious, and wary. The term behavioral inhibition has been used to characterize individual differences in children's reactions to novel social and nonsocial situations (García Coll, Kagan, & Reznick, 1984; Schmidt & Buss, 2010). Whereas inhibition in infants is assessed mainly in terms of their latency to approach novel objects or nonsocial stimuli, research on inhibition in older children often focuses on their reluctance to interact spontaneously with unfamiliar adults or peers (e.g., Asendorpf, 1991). Inhibition of approach to unfamiliar peer situations is typically characterized by onlooker, unoccupied, and other anxious behaviors (e.g., watching others playing without joining in, whining, nail biting) (e.g., Asendorpf, 1991; Rubin, Coplan, Fox, & Calkins, 1995). From a different perspective, researchers who study shyness as a personality trait, especially in adults, focus on feelings of self-consciousness and anxiety in social interactions (e.g., Cheek & Buss, 1981). Asendorpf (1991) characterized shyness as deriving from an internal conflict of approach and avoidance motivations in social settings. Shy children are interested in social interactions, but this approach motivation is hindered by fear and wariness. Shyness and inhibition both seem to tap individual reactivity to challenging situations, with the former focusing on the anxious response to social novelty and the latter focusing on the dispositional characteristic to be fearful when encountering unfamiliar social or nonsocial situations (Coplan & Armer, 2007; Rubin, Coplan, & Bowker, 2009; Schmidt & Buss, 2010).

Considering the conceptual and empirical overlap between shyness and inhibition and their similar behavioral manifestations, X. Chen and French (2008) used the term shynessinhibition to refer to children's wary and anxious reactivity to stressful or challenging social situations. Shyness-inhibition is similar to reticence, anxious solitude, shyness-sensitivity, and some other constructs that researchers use to describe behaviors that indicate internal fearfulness in social settings (e.g., Coplan & Armer, 2007; Gazelle & Ladd, 2003). However, it is different from those constructs representing social withdrawal, such as unsociability or social disinterest due to the lack of desire to interact with others (e.g., “would rather be alone”) (Coplan & Armer, 2007). Distinction between shyness-inhibition and unsociability or social disinterest is important in cross-cultural research because they may be valued differently in different cultures. For example, in Western individualistic societies, relative to shyness-inhibition, unsociability that is driven by preference for solitude may be viewed as more normal because it is sometimes considered indicative of personal choice and conducive to performance on constructive tasks and emotional health (e.g., Burger, 1995; Coplan et al., 2004; Leary, Herbst, & McCrary, 2003). In contrast, in some group-oriented cultures, shynessinhibition may be viewed as acceptable or even desirable, but unsociability is regarded as anticollective and abnormal, a risk factor in development (Casiglia, Lo Coco, & Zappulla, 1998; X. Chen, Wang, & Cao, 2011; Valdivia, Schneider, Chavez, & Chen, 2005). A major aspect of socioemotional functioning that largely overlaps with shyness-inhibition is social anxiety. The construct of social anxiety taps fear, concern, and worry about negative social evaluations and distress in familiar and unfamiliar social situations (American Psychiatric Association, 1994). Whereas shyness-inhibition is mainly concerned with behavioral manifestations of internal conflict or fearful feelings, social anxiety focuses on emotional aspects of individual reactions to challenging social situations. Shyness-inhibition has traditionally been conceptualized from temperament or personality and peer relationship perspectives, whereas the experience of social anxiety is often discussed in the framework of psychopathology. In addition, researchers use various methods, including observations, peer evaluations, and self-reports, in the assessment of shy-inhibited behavior, but research on the emotional experience of social anxiety relies mostly on self-reports and clinical assessments. Nevertheless, the literature has indicated robust links between shyness-inhibition and social anxiety in children and adolescents, both having considerable implications for peer interactions and relationships (e.g., Rubin et al., 2009; West & Newman, 2007). As another peer-relevant aspect of socioemotional functioning, loneliness is associated with both shyness-inhibition and social anxiety (e.g., Rotenberg & Hymel, 1999; Rubin et al., 2009). Loneliness typically refers to the subjective experience of dissatisfaction with one's socialrelational life and related emotional reactions of sadness and emptiness (e.g., Asher & Paquette, 2003). Research indicates that children as young as preschool age report loneliness and social dissatisfaction, and loneliness in childhood is associated with unpleasant feelings and perceptions of unfulfilled relational needs such as lack of companionship and support (e.g., Cassidy & Asher, 1992; Coplan, Closson, & Arbeau, 2007). Moreover, like shynessinhibition and social anxiety, feelings of loneliness are stable over time and may predict adjustment problems, such as depression, suicide, and psychosomatic problems (e.g., Aanes,

Mittelmark, & Ketland, 2010; Qualter, Brown, Munn, & Rotenberg, 2010). Shyness-Inhibition Across Cultures One of the primary issues in cross-cultural research is whether children in different societies engage in different levels or types of social interactions (e.g., Whiting & Edwards 1988). Edwards (2000), for example, found that children in relatively close and rural societies (e.g., Kenya and India) had significantly lower scores on overall social engagement than children in more open and urban societies (e.g., Okinawa and the United States) where peer interactions were encouraged. Researchers also found that, compared with their North American counterparts, children in many non-Western rural societies, such as Bedouin Arab, Kenya, Maya, Mexico, and India (Ariel & Sever, 1980; Edwards, 2000; Farver & Howes, 1993; Gaskins, 2006), tended to be less expressive of their personal styles in peer interactions and engaged in little sociodramatic activities that required control of social-evaluative anxiety. These later results are particularly interesting because the lack of self-expression and assertiveness in social interactions is closely related to shyness-inhibition. Researchers consistently found that East Asian children were more shy-inhibited than North American children in social settings (e.g., X. Chen, 2010; Farver & Howes, 1988). In a study of play behavior in Korean preschools in the United States, for example, Farver, Kim, and Lee (1995) found that, compared with European American children, Korean American children displayed more shy behaviors. Moreover, Korean children's play included less self-expressive themes and Korean children used less self-assertive communicative strategies (Farver & Shin, 1997). Similarly, Rubin, Hemphill, et al. (2006) found that Korean and Chinese children exhibited higher inhibition and anxiety than Australian, Canadian, and Italian children in the laboratory situation, with the former displaying more vigilant and reactive behaviors such as staying closer to their mothers and being reluctant to explore in free play sessions or in interactions with a stranger. When the child was asked by the experimenter to touch a scary toy robot, the mean latency to touch the toy was 18.97 and 13.97 seconds in Korean and Chinese children but was 7.74, 7.70, and 7